Sie sind auf Seite 1von 161

ABSTRACT

HUTCHINSON, SHAWN ROBERT. Thermoplastic polyacrylonitrile: investigation

of polymer structure, melt behavior and fiber properties. (Under the direction of

W. OXENHAM, B.S. GUPTA and A.E. TONELLI).

The objective of this research is to understand the structure, melt

behavior, and fiber properties of Amlon® D, part of a family of thermoplastic high

acrylonitrile (PAN) copolymer resins. A proposed crystal morphology model

based on predicted and observed comonomer sequence lengths and melting

behavior is presented. Thermograms of basic melt behavior demonstrate there

is melting, but surprisingly no evidence of a first-order phase transition is

observed.

Time-temperature stability and flow behavior was used to gain an

understanding of processing parameters. Thermogravimetric analysis with mass

spectrometry was used to study the stability. A capillary rheometer was used to

obtain detailed rheology data employing a range of die sizes, melt times and

shear rates.

Monofilaments were extruded using a single bore device with a range of

spin draw and draw ratios. Fiber tenacity and elongation were measured to

assess the efficacy of processing conditions. The solid state behavior of the

polymer is characterized and correlated with the mechanical properties using

wide angle x-ray diffraction and differential scanning calorimetry analyses.


THERMOPLASTIC POLYACRYLONITRILE

by

SHAWN ROBERT HUTCHINSON

A thesis submitted to the Graduate Faculty of

North Carolina State University

In partial fulfillment of the

Requirements for the degree of

Master of Science

TEXTILES

COLLEGE OF TEXTILES

Raleigh, NC

2005

Approved by:

_____________________________ _____________________________

_____________________________ _____________________________
Chairman of the Advisory Committee Co-Chairman of the Advisory Committee
ii

DEDICATION

This work is dedicated to

my parents,

sister,

family and

friends,

whose support made with work possible.

Thermoplastic Polyacrylonitriles
iii

THE ROAD NOT TAKEN

By Robert Frost

TWO roads diverged in a yellow wood,

And sorry I could not travel both

And be one traveler, long I stood

And looked down one as far as I could

To where it bent in the undergrowth; 5

Then took the other, as just as fair,

And having perhaps the better claim,

Because it was grassy and wanted wear;

Though as for that the passing there

Had worn them really about the same, 10

And both that morning equally lay

In leaves no step had trodden black.

Oh, I kept the first for another day!

Yet knowing how way leads on to way,

I doubted if I should ever come back. 15

I shall be telling this with a sigh

Somewhere ages and ages hence:

Two roads diverged in a wood, and I—

I took the one less traveled by,

And that has made all the difference. 20

Thermoplastic Polyacrylonitrile
iv

AUTOBIOGRAPHY

Shawn Robert Hutchinson was born on February 2, 1978 in Frankfurt am

Main, Hessen, Germany. Childhood was spent along the banks of the Ohio

River in Vienna, West Virginia collecting Native American artifacts and cultivating

towering gardens. In 1996, he graduated from Fuquay-Varina Senior High

School Salutatorian and Student Body President. After graduating cum laude

from Duke University with a B.A. in Computer Science and German minor, he

worked as a software developer with a dotcom in Washington, D.C. and New

York, NY.

Seeking to fulfill questions about the nature of existence, he sought an

experiential worship that was rediscovered in creativity – ultimately finding

Vedanta and Santhana Dharma as a profound system of self-knowledge. After a

period with the Sivananda Yoga Vedanta Center and travels in India, he returned

to graduate school at North Carolina State University to complete the

development of a new fabric. Unbeknownst to him lie waiting the world of

polymers and fibers. How thankful he is to have these human experiences and

hopes to utilize his skills most effectively in improving the human condition.

Thermoplastic Polyacrylonitrile
v

ACKNOWLEDGEMENTS

The author would like to express his sincere appreciation to the committee

chairpersons Drs. W. Oxenham and B.S. Gupta. Their support, motivation and

suggestions made this project possible. Dr. Gupta, especially, provided indelible

guidance on the ethics of scientific inquiry. Candid debates with Dr. A.E. Tonelli

provided insight into the structure and properties of polymers and inspiration for

an aspiring scientist. Appreciation is also extended to members Drs. J.P.

Hinestroza and D.A. Shiffler. Many discussions with technical advisor C. Moses,

of the Institute of Textile Technology provided enduring encouragement and

support. Without hesitation, Dr. K.R. Beck invested his time into teaching

analytical techniques and skills. Drs. S.M. Hudson and R. Kotek provided key

recommendations along the journey.

Eine hertzliche Bedanken für Birgit Anderson and Ted Dodson, whose

technical assistance made this project possible. Jason Dinsmore and Sabapathy

Sankar of the Chemistry Department at North Carolina State University provided

nuclear magnetic resonance training. Tulane University, Drs. D. De Kee, Y.

Zhong, W. Liu, and the Tulane Institute for Macromolecular Engineering and

Science (TIMES) graciously allowed their equipment to be used for the

rheometry and part of the thermogravimetric analysis work. Thanks to Dralon,

GmbH for allowing their fibers to be the subject of investigation free of charge.

This work was sponsored by the Master’s Graduate Fellowship Research

Program from the Institute of Textile Technology. The financial support of the

member companies is greatly appreciated.

Thermoplastic Polyacrylonitrile
vi

TABLE OF CONTENTS

LIST OF TABLES............................................................................................... IX
LIST OF FIGURES .............................................................................................. X
0 INTRODUCTION ...............................................................................................1
1 ACRYLIC CHEMISTRY.....................................................................................3
1.1 ACRYLIC MONOMERS......................................................................................4
1.2 MONOMER SYNTHESIS ...................................................................................5
1.3 POLYMERIZATION ...........................................................................................7
1.3.1 Bulk Polymerization .........................................................................7
1.3.2 Solution Polymerization ...................................................................8
1.3.3 Heterogeneous (Emulsion & Suspension) Polymerization ..............9
2 POLYACRYLONITRILE FIBER TECHNOLOGY ............................................10
2.1 HIGH ACRYLICS & MODACRYLICS .................................................................10
2.2 BICOMPONENT FIBERS & CRIMP ...................................................................12
2.3 DYEABILITY .................................................................................................14
2.3.1 Acid & Base Dyes..........................................................................14
2.3.2 Producer Dyed ..............................................................................14
2.3.3 Pigmented .....................................................................................15
2.4 TECHNICAL APPLICATIONS ...........................................................................16
2.4.1 Reinforcing Fiber (Asbestos Replacement) ...................................16
2.4.2 Carbon Fibers................................................................................16
2.4.3 Moisture Absorbent Fibers ............................................................17
3 FIBER SPINNING TECHNOLOGY..................................................................18
3.1 WET SPINNING ............................................................................................19
3.1.1 Process .........................................................................................19
3.1.2 Solvents.........................................................................................19
3.1.3 Polymer .........................................................................................20
3.1.4 Coagulation ...................................................................................20
3.1.5 After treatment...............................................................................21
3.2 DRY SPINNING .............................................................................................23
3.2.1 Process .........................................................................................23
3.2.2 Solvents.........................................................................................24
3.2.3 Polymer .........................................................................................24
3.2.4 Evaporation & Cross Sectional Shape...........................................25
3.2.5 After treatment...............................................................................25
3.3 MELT SPINNING ...........................................................................................27
3.3.1 Conventional Thermoplastics Processing......................................27
3.3.2 Imparted melt processability..........................................................28
3.3.3 PAN Hydrates................................................................................29
3.3.4 Summary of Prior Art in ‘Melt Processable’ PANs .........................29

Thermoplastic Polyacrylonitrile
vii

4 MARKETS AND POTENTIAL .........................................................................30


4.1 TRENDS IN THE WORLDWIDE SYNTHETIC FIBER PRODUCTION .........................30
4.2 TRENDS IN THE ACRYLIC FIBER MARKET .......................................................30
4.2.1 Markets in the United States .........................................................32
4.2.2 Markets in Europe and Asia ..........................................................39
4.2.3 Market Prospects...........................................................................39
4.3 COST ANALYSES .........................................................................................40
4.3.1 Poly(ethylene terephthalate)..........................................................40
4.3.2 Poly(propylene) .............................................................................41
4.3.3 Poly(amides) .................................................................................41
4.3.4 Solution processable high acrylonitrile copolymers .......................42
4.3.5 Melt processable high acrylonitrile copolymers .............................43
4.4 POTENTIAL APPLICATIONS FOR THERMOPLASTIC PAN ....................................48
4.4.1 Comparative properties between synthetic fibers..........................48
4.4.2 Deniers, profiles and multicomponent fibers..................................49
4.4.3 Reinforcing fiber ............................................................................50
4.4.4 Carbon fibers.................................................................................50
4.4.5 Composite and carbon carbon composite components.................52
4.4.6 Limitations from the hot-wet conundrum........................................53
4.4.7 Outdoor fabrics ..............................................................................54
4.4.8 Nonwovens....................................................................................54
4.4.9 Membranes, Films and Conductivities...........................................54
5 POLYMER MICROSTRUCTURE ....................................................................57
5.1 EXPERIMENTAL ............................................................................................61
5.1.1 Materials........................................................................................61
5.1.2 Solution 13C Nuclear Magnetic Resonance ...................................62
5.1.3 Differential Scanning Calorimetry ..................................................63
5.2 RESULTS AND DISCUSSION ...........................................................................63
5.2.1 Solution 13C Nuclear Magnetic Resonance ...................................63
5.2.2 Differential Scanning Calorimetry ..................................................70
5.2.3 Conclusions...................................................................................72
6 POLYMER THERMOMECHANICAL ANALYSES ..........................................76
6.1 EXPERIMENTAL ............................................................................................77
6.1.1 Drying ............................................................................................77
6.1.2 Fourier Transform Infrared Spectroscopy......................................77
6.1.3 Thermal Gravimetric Analysis & Mass Spectrometry.....................78
6.1.4 Capillary Rheometry ......................................................................78
6.2 RESULTS AND DISCUSSION ...........................................................................81
6.2.1 Vacuum Drying ..............................................................................81
6.2.2 Chemical Composition...................................................................81
6.2.3 Thermal Stability............................................................................83
6.2.4 Degradation Byproducts ................................................................85
6.2.5 Melt Time.......................................................................................86
6.2.6 Capillary Rheometry ......................................................................90

Thermoplastic Polyacrylonitrile
viii

6.2.7 Conclusions...................................................................................91
7 MELT EXTRUSION .....................................................................................95
7.1 EXPERIMENTAL .......................................................................................98
7.1.1 Extrusion Device .........................................................................98
7.1.2 Mechanical Properties .................................................................102
7.1.3 Size Exclusion Chromatography .................................................102
7.1.4 Wide Angle X-ray Diffraction .......................................................102
7.1.5 Differential Scanning Calorimetry ................................................103
7.2 RESULTS & DISCUSSION ............................................................................103
7.2.1 Melt Spinning...............................................................................103
7.2.2 Mechanical Properties .................................................................105
7.2.3 Molecular Weight.........................................................................107
7.2.4 Wide Angle X-ray Diffraction .......................................................119
7.2.4.1 Background ...............................................................................119
7.2.4.2 Results and Discussion .............................................................121
7.2.5 Paracrystallinity ...........................................................................127
7.2.6 Conclusions.................................................................................131
8 GENERAL CONCLUSIONS ..........................................................................132
8.1 FUTURE WORK ..........................................................................................133
9 LITERATURE CITED ....................................................................................134

Thermoplastic Polyacrylonitrile
ix

LIST OF TABLES

Table 1. Raw Materials and Conversion Costs for Major Thermoplastic Polymers

.....................................................................................................................44

Table 2. Comparative Properties of Synthetic Fibers..........................................49

Table 3. Comonomer Distributions and Number Average Sequence Lengths for

Methyl Acrylate and Vinyl Acetate Copolymers ...........................................67

Table 4. Thermogram Inflection Points and Enthalpies for Dried and Undried

Amlon® D Pellets .........................................................................................73

Table 5. Capillary Rheometry: Die Dimensions and Shear Rates......................79

Table 6. Extrusion Conditions: Melt Temperature and Melt Times ....................80

Table 7. Power Law Index for Dried Amlon® D Capillary Rheometry ..................90

Table 8. Melt Extrusion, Winding and Drawing Conditions ...............................100

Table 9. Tenacity and Elongation for Meltspun Amlon® D Fibers by Total

Effective Draw Ratio ..................................................................................104

Table 10. Size Exclusion Chromatography Retention Times for Amlon® D Fibers

...................................................................................................................108

Table 11. Wide Angle X-Ray Diffraction Reflection Spectrum Analysis of Amlon®

D Fibers .....................................................................................................123

Table 12. Differential Scanning Calorimetry Data for Amlon® D Fiber ..............128

Thermoplastic Polyacrylonitrile
x

LIST OF FIGURES

Figure 1. Poly(acrylonitrile-co-methyl acrylate) ...................................................11

Figure 2. Synthetic Fibers Market: Acrylics, Polyester, .......................................31

Figure 3. Capacity, Production, Domestic Shipments, Imports, Exports and

Consumption in United States Acrylic Fiber Market.....................................33

Figure 4. Capacity, Production and Consumption in the Acrylic Fiber Markets by

Major Geographical Regions .......................................................................34

Figure 5. Acrylic Fiber Consumption by Apparel, Home .....................................36

Figure 6. Synthetic Fibers Market: Acrylics, Polyester, .......................................38

Figure 7. Raw Material, Intermediary and Fiber Costs for Poly(ethylene

terephthalate), Poly(propylene), Poly(amides) and Polyacrylonitriles. .........46

Figure 8. Solution 13C NMR Spectra of Amlon® D in d6-DMSO...........................68

Figure 9. Number Average Acrylonitrile Sequence Length for Methyl Acrylate

Copolymers by Mole Fraction of Acrylonitrile: Observed Lengths and Zero-

Order, First-Order, and Second-Order Markov Models Lengths ..................69

Figure 10. Differential Scanning Calorimetry Cooling Thermograms for

Poly(ethylene terephthalate) Tire Cord Fibers .............................................72

Figure 11. Differential Scanning Calorimetry Thermograms for ..........................74

Figure 12. Weight Loss during Vacuum Drying of Amlon® D Pellets...................82

Figure 13. Fourier Transform Infrared Spectra of Amlon® D Pellets ...................82

Figure 14. Quasi-Isothermal Stability Thermogravimetric Analyses....................84

Figure 15. Isorate Degradation Thermogravimetric Analyses .............................84

Thermoplastic Polyacrylonitrile
xi

Figure 16. Mass Spectrometry Fragments of Undried Amlon® D Pellets Collected

with Tandem Isothermal Thermogravimetric Analyses. ...............................87

Figure 17. Melt fractures of Dried Amlon® D Extrudate.......................................92

Figure 18. True Viscosities by True Shear Rate for Dried and Undried Amlon® D

and Polypropylene by Temperature and Melt Time .....................................93

Figure 19. Melt Extrusion Device with Detail of Bore and Shroud Setup ..........101

Figure 20. Linear Density (denier) of Melt-spun Amlon® D Fiber ......................109

Figure 21. Load by Elongation Curves for Amlon® D Fiber, Plastic Deformation

and Deformation Strengthening .................................................................110

Figure 22. Tenacity for Amlon® D Fiber by Total Effective Draw Ratio .............111

Figure 23. Elongations for Amlon® D Fiber .......................................................112

Figure 24. Tenacity by Elongation and Spin Draw Ratio for Amlon® D Fiber ....113

Figure 25. Tenacities by Draw Ratio for Amlon® D Fiber ..................................114

Figure 26. Elongations at Break for Amlon® D Fiber .........................................116

Figure 27. Elongation at Break for Amlon® D Fiber...........................................117

Figure 28. Wide Angle X-Ray Diffraction Reflection Spectrum for Fibers Extruded

with a Heat Shroud. ...................................................................................124

Figure 29. Wide Angle X-Ray Diffraction Transmission Patterns for Fibers

Extruded with a Heat Shroud .....................................................................125

Figure 30. Second Cooling Enthalpies for Amlon® D Fiber ...............................129

Thermoplastic Polyacrylonitrile
1

0 Introduction

The ability to precisely control the physical structure of materials

profoundly affects their properties. This ability will be one of the defining features

of the new century. Acrylic is a unique example.

The commercial history of polyacrylonitrile dates back to the 1950s when

Bayer and DuPont patented dry spinning technologies. The polymer degraded

before melting, which rendered it processable only in harsh, toxic solutions.

Inherent ultraviolet (UV) and chemical resistances made the fibers special in a

number of applications and, combined with its high resiliency, a fiber of choice as

a wool substitute at that time. However, low throughputs and expensive solvent

recovery associated with solution processing made the fibers expensive.

Variations in fiber diameter and surface imparted from solution processing, as

well as competition from nylon, limited the success of the fiber. Lack of melt

processability prevented the polymer from reaching its full potential in all

contemporary thermoplastic polymer applications.

For decades, scientists sought ways to melt-process polyacrylonitrile,

because of low conversion costs and potential mechanical property

enhancements. Numerous attempts were made to impart melt-processability,

but few, if any, met commercial success. Finally, in the 1990s, British Petroleum

(BP) developed and patented a polymerization reaction that created a family of

the first innately thermoplastic high acrylonitrile copolymers under the trade name

Amlon®. BP claims that under extremely low feed rates, the composition of a

copolymer can be controlled sufficiently to impart melt processability. Thus,

Thermoplastic Polyacrylonitrile
2

acrylic can now be processed in an environmentally friendly manner without the

aid of solvents. However, little is publicly known about how to process this

polymer. Generating such data and investigating the structure-property

relationship of resulting fibers is essentially the focus of this study.

The objective of this thesis is to first present an economic picture of the

U.S. and worldwide acrylic fiber market, synthetic fiber market, and raw material

costs. Potential applications including fibers, films, fabrics and composites are

explored. Secondly, the basic polymer science in terms of copolymer structure

and crystal morphology is investigated to understand how polyacrylonitrile can be

made melt processable. Next, melt behavior, stability and rheology are studied

to determine optimal processing conditions. Finally, fibers were melt-extruded

and correlations were made between extrusion conditions, microstructure and

mechanical properties.

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 3

1 Acrylic Chemistry

Vinyl monomers comprise a vast majority of modern synthetic chemicals.

Acrylic is a constituent of a well-known subset of Acrylic Butadeine Styrene

(ABS) plastics. ABS materials include some of the most specialized and highly

valued of all petrochemicals. In particular, their chemical, thermal and ultraviolet

(UV) resistances enable them to satisfy a technical niche all their own. Even

amongst apparel markets, acrylonitrile provides an important alternative to

woolens. ABS polymer applications affect nearly every level of daily human

existence.

Polyacrylonitrile (PAN) and its copolymers are the only ABS polymers

used today in fiber applications because it is sufficiently paracrystalline, which

gives it the mechanical stability necessary for applications. The acrylonitrile

sidegroups impart unique resistances and the strong interactions between them

traditionally prevent the material from melting before it degrades. At

temperatures above 200ºC, chemical changes begin to occur as the sidegroups

begin to crosslink and cyclize. On one hand, PAN has desirable properties that

are not found in other commodity fiber forming polymers such as polyamide,

polypropylene and polyethylene terephthalate. However, these properties also

prevent the polymer from being used in more general consumer plastic

applications such as films because of processing requirements.

In the 1950s companies like DuPont and Bayer developed the first

specialized systems to process PAN in solution. Yet, the environmental and

economic constraints from harsh solvents and low throughputs limited the

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 4

commercial success of PAN applications. The nature of the required processing

equipment itself restricted most of the uses to fibers. Accordingly, many

scientists and chemical companies attempted to alter the chemical behavior so

as to impart melt processability. To date, however, none of the techniques have

proven successful on a mass scale due to pervading deficiencies in the

mechanical properties of products and the economics of their production.

1.1 Acrylic monomers

‘Acrylic’ is the colloquial term for polyacrylonitrile, or PAN, and its

copolymers. However, in organic chemistry, ‘acrylic’ applies to a classification of

vinyl monomers having the basic form, where R1 and R2 are sidegroups:

H R1
C C
H C O
O R2

The most widely used forms of acrylics are:

• CH2=CH⎯COOH, arylic acid

• CH2=CHCH3⎯COOH, methacrylic acid

• CH2=CH⎯COO⎯CH3, methyl acrylate

• CH2=CCH3⎯COO⎯CH3, methyl methacrylate (Plexiglas®)

• CH2=CH⎯CO⎯NH2, acrylamide

• CH2=CH⎯CHO, acrolein

• CH2=CH⎯CN, acrylonitrile (AN)

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 5

Other side groups in place of the R1 and R2, such as halides, enable a variety of

property enhancements including flame resistance and dyeability.

In particular, methyl methacrylate is one of the mostly commonly used

acrylic monomers due to their aging resistance. With a relatively easy bulk

polymerization process, this monomer lends itself well to molding and casting at

low temperatures, from 45 to 90 °C. In bulk, the liquid polymer requires curing

from 1.25 to 2.5 hours ranging in temperature from 250 to 315 °C.

Cast methyl methacrylate is available in a variety of shapes – generally

sheets, rods and tubes. The properties of these thermoplastic acrylics come in

three grades: 1) outdoor use where moderate transparency is required having

half the weight and tenfold impact resistance of glass, 2) heat-resistant that

resists crazing and solvents, and 3) technical applications where high optical

clarity and low defects are required. Cast materials are available in a variety of

sizes with auxiliary properties from colors to scratch resistance [2].

1.2 Monomer Synthesis

Most acrylic monomers are synthesized through an initial reaction of

hydrogen cyanide (HCN). Originally, methyl methacrylate was synthesized with

acetone and hydrogen cyanide. Later, ethylene oxide replaced acetone. Other

processes that form higher acrylates are alcoholysis or transesterification.

Catalysts such as toluene-sulfonic acid are used to increase yields. Inhibitors

such as hydroquinone are often used to prevent spontaneous polymerization.

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 6

Aside from being the raw material for the polymerization of PAN,

acrylonitrile is often also used as a precursor for high acrylic esters. The three

main means of synthesizing acrylonitrile are 1) from propylene, ammonia and

oxygen, 2) the dehydration of ethylene cyanohydrin into acrylonitrile and water

with a catalyst and 3) a reaction of acetylene and HCN at high temperature with a

catalyst, usually nickel carbonyl and hydrochloric acid:

1) 4CH2=C=CH2 + 4NH3 + 5O2 ⎯⎯⎯→ 4CH2=CH⎯CN + 10H2O


catalyst

2) OH⎯CH2⎯CH2⎯CN ⎯⎯⎯→ CH2=CH⎯CN + H20


catalyst

400−500o C
3) C2H2 + HCN ⎯⎯⎯⎯→ CH2=CH⎯CN
catalyst

The hydrolysis of acrylonitrile and methacrylonitrile is also used to synthesize

acrylic acid and methacrylic acid:

H+
CH2=CH⎯CN + H2O ⎯⎯→ CH2=CH⎯COOH + NH3

In general, most acrylic monomers are synthesized with a catalyst directly from

HCN or indirectly through acrylonitrile. Regardless, the reactions involve

explosive combinations of high temperatures, high pressures and large volumes

of cyanides. Thus, extreme care is required when handling the reactions and

monomers [2].

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 7

1.3 Polymerization

The two sidegroups, R1 and R2, provide the means for a wide array of

constituent compounds. Polymerization occurs between the double-bonded

carbon atoms and creates high molecular weight polymers. The degree of

polymerization and average molecular weight can create significant phase

differences at room temperatures ranging from an oily liquid to a brittle solid.

Most acrylic bulk polymers are thermoplastic, or they soften at a specific

temperature. Thermosetting, or crosslinking polymer chains, can decrease and

altogether vanquish the plasticity of the polymer. Dimethacrylates and diallylic

compounds are often used as thermosetting agents. Partially crosslinked

polymers can yield desirable properties such as better heat resistance while

retaining the plastic nature [2].

As in the synthesis of the monomer, polymerization is often a highly

exothermic reaction. Control of the reaction and temperature are highly coupled.

Polymerization of acrylics usually occurs in three ways: 1) bulk, 2) solution and

3) heterogeneous systems.

1.3.1 Bulk Polymerization

Bulk polymerization takes place directly on the monomer with the aid of a

soluble catalyst and is used for most cast materials. The polymer melt is allowed

to pre-polymerize to about 10% before it is poured into the mold. The mold is

subjected to low temperatures, allowing the reaction to continue until the rate

significantly decreases. At this point, the temperature is increased to the boiling

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 8

point of the monomer to complete the reaction. This method is used primarily for

2-dimensional molds. Autoclaves are used for bulkier dimensions. The high

molecular weight products are rigid and resistant to solvents.

1.3.2 Solution Polymerization

Lower molecular weight polymers are synthesized through solution

polymerization. In this variation, the monomer and catalyst are added to a

solvent in which both the monomer and polymer are soluble. Ethyl acetate,

toluene and acetone are favorable solvents. As a result, this method is useful

when the end product contains both solvents and the acrylic polymer, such as

lacquer, some adhesives and most water-resistant films. The molecular weight is

controlled by the amount of catalyst, ratio of solvent to monomer and reaction

temperature [2].

Solution polymerization is used by many of the PAN fiber producers as a

means of synthesizing the polymer in a form that is easily solution-spun because

the polymer is already in a spinning solution. Materials having as little as 50%

acrylonitrile by weight can be synthesize by this method. For vinyl compounds,

dimethylformamide is usually the solvent. A range of patents to various chemical

companies in the 1960s-1970s outlined a variety of initiators, catalysts, and

processes [3].

Thermoplastic Polyacrylonitrile
Acrylic Chemistry 9

1.3.3 Heterogeneous (Emulsion & Suspension) Polymerization

Of the heterogeneous polymerization mechanisms, emulsion and

suspension occupy the vast majority of preferred methods. For emulsion, an

emulsifier and catalyst are dissolved in water. The monomer and surfactant are

added to this mixture and agitated. A variety of emulsifiers and surfactants yield

myriad reactions. The surfactant is arguably the most important part of the

system as it is responsible, at sequential stages, first as a dispersant, second as

a protective material to prevent coagulation and finally to suspend the solid

polymer particles in water for a stable mixture.

Suspension polymerization reacts in a similar way to emulsion, yet the

monomer is dispersed into small droplets. A monomer-soluble catalyst is then

added. Thus, unlike the emulsion system where polymerization occurs in water,

suspension polymerization occurs in the monomer droplets. After the reaction is

complete, the water is drained and droplets are dried for further processing [2].

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 10

2 Polyacrylonitrile Fiber Technology

In the 1950s, DuPont created the first acrylonitrile fiber, Orlon®. It was a

dry-spun fiber composed of 100% acrylonitrile. This fiber had several defects

that led to the next stage in polymer engineering. In particular, the high polarity

of the acrylonitrile sidegroups and rigid conformation made solution processing

difficult. Post-spinning processes such as dyeing were also adversely affected.

As a result, copolymerization with vinyl or vinylidene compounds enabled the

structure to be opened and softened [4].

2.1 High Acrylics & Modacrylics

The degree of copolymer delineates two types of acrylonitrile compounds:

1) high acrylic fibers consisting of at least 85% acrylonitrile (AN) and 2) modified

acrylics (modacrylics) consisting of 35-85% acrylonitrile. Typical high acrylonitrile

comonomers include methyl methacrylate, vinyl acetate or methyl acrylate, the

latter is shown in Figure 1.

Flammability and ignition behavior of high acrylic and modacrylics are

similar to other synthetic fibers like nylon, polyester, olefins and wool. They tend

to melt and burn rather than char and burn as do cellulose fibers. Of the

synthetics, however, high acrylic has a Limiting Oxygen Index similar to cotton.

Both are considered relatively flammable. For applications where flammability is

key, such as carpets, sleepwear, draperies and bedding, fibers must be self-

extinguishing [5]. To pass flammability tests, halogen compounds such as vinyl

chloride, vinylidene chloride and vinyl bromide are used. Commercially, only the

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 11

halogen modacrylics provide sufficient flame resistance for any application other

than carpet. Nearly all flame retardant high acrylic production has ceased [6].

Due to the low boiling points of halogens, polymerization of modacrylics is

conducted under increased pressures. Vinyl chloride (VC) requires the highest

pressure and is relatively unreactive in copolymerization with acrylonitrile. Free

radical polymerization of polyvinyl chloride and acrylonitrile occurs relatively fast.

Copolymerization by means of adding an AN monomer to a VC-terminated chain

occurs quickly. However, adding a VC monomer to an acrylonitrile-terminated

chain occurs slowly. Thus, the concentration of AN is low. The free radical is

unlikely to react with the vinyl chloride monomer. Thus, the copolymerization

process requires excess VC and is controlled by the concentration of AN

monomer. This reaction is generally conducted in a batch process. Molecular

weight and separation of toxic unreacted monomer presents difficulties.

* CH2 CH n CH2 CH m *
C C O
O
N
CH3

Figure 1. Poly(acrylonitrile-co-methyl acrylate)

Unlike VC, copolymerization of vinylidene chloride is conducted at lower

pressures with a more balanced concentration to inclusion ratio. Vinyl bromide

has a similar reaction rate and reaction mechanism. However, bromide is

seldom used solely as a flame retardant. Generally, to create copolymers of

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 12

sufficient molecular weight for fibers, the reaction rate is limited by a required low

concentration of initiator [6].

Most modacrylics are wet-spun using solvents suitable for acrylonitrile.

Dynel® appears to be the only exception to this process as it is either dry-spun

with acetone or melt spun. During spinning, other compounds are added for

increased flame retardance (antimony oxides), light stabilizers (titanium dioxide)

and color stabilizers (pigments).

Wet spinning of modacrylics is conducted by a similar method to high

acrylics. However, as halogens are sensitive to heat, minimizing heat exposure

is critical during solution- and fiber-formation stages. During processing, small

amounts of HCl and HBr are evolved creating mildly corrosive acids. Either in

spinning or in customer processes, a weak base or epoxide is often added after

heating to reduce acidity.

Other comonomers are used to alter the behavior of acrylics. Acrylamide

is often used to control shrinkage and physical properties. Sulfonates improve

dyeability [6].

2.2 Bicomponent Fibers & Crimp

The class of bicomponent fibers come in many forms and can be spun

from either the melt or solution. To date, the only commercially available acrylic

bicomponent is a side-by-side composite, developed by DuPont. The

construction, due to the composition of the two components creates a crimp

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 13

similar to that observed in wool from the ortho- and para-cortexes. Belonging to

this group of crimped bicomponents are water-reversibles and irreversibles.

In 1958, DuPont introduced the only water-reversible (through drying)

Type 21 Orlon® bicomponent using hydrophobic PAN and a hydrophilic

copolymer of AN with a sulfonate monomer. Although the spinerette holes are

circular, a mushroom-shaped cross section is formed from forces in the gellation

process. PAN is the ‘cap’ and AN/sulfonate is the ‘stem.’ Type 21 Orlon® as

sold to the textile mill has primarily mechanical crimp. However, through

exposure to water in dyeing and drying, a spiral crimp can be developed. The

hydrophilic stem shrinks as water is removed. Uniquely, the properties of the

AN/sulfonate compound enable this swelling and contraction to occur repeatedly.

However, the mechanical forces providing the crimp are relatively low, so the

crimp response only occurs when the fabric is not hung or blocked.

Irreversible crimps can be constructed using a variety of copolymers. The

two types of copolymers can both contain high contents of acrylonitrile, a high

and modacrylic levels, or both modacrylics. Discontinued in 1990, DuPont’s

Type 24 Orlon® was an example of a dry-spun irreversible crimped bicomponent.

Wet-spun bicomponents are more difficult to assemble. Only two technologies

have emerged with various feed patterns. In either case, ensuring a consistent

cross section composition is difficult – ranging from 100% component A to 100%

component B with a sandwich construction in between. Non-water-reversible

bicomponent crimps are often used in blends with a monocomponent fiber where

bulk and hand are optimized [6].

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 14

2.3 Dyeability

2.3.1 Acid & Base Dyes

Dyeing acrylic with an acid was the original method of coloring fibers.

There were a number of drawbacks that prevented this from becoming the de

facto standard. First, the dye sites were based on weak bases such as 2-vinyl

pyridine. A strong acid was thus required in the bath to protonate the site and

make it dyeable. Second, the amine group reduced the heat stability of the fiber.

Further, the base color of the acid-dyeable fiber was cream and the fastness was

not noteworthy. Finally, cost of the monomer was prohibitive.

The introduction of basic (cationic) dyes by DuPont and Bayer in the

1950s, Sevron® and Astrazon® dyes respectively, relegated acid-dyeable fibers

to a specialty niche. The range of dyes was expanded as they had strong affinity

for the sulfonic and sulfatic dyesites introduced from the polymerization initiator.

Basic-dyeable fibers were less expensive as the monomer could be easily

recovered in distillation enabling a continuous dyeing range. Further, these

fibers had a whiter base and could be dyed at a more neutral pH with brilliant

colors having better fastness [6].

2.3.2 Producer Dyed

When considering total worldwide acrylic fiber production, producer-dyed

fibers occupy a small market share. However, in their application, they offer

substantial cost savings for more commodity colors. Dyeing in the production

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 15

stage occurs by three means: 1) gel state, 2) solution (dope) and 3) tow. During

wet spinning, acrylic fibers have a porous structure of at least 50% void volume

that is well suited for dyeing. Dye penetration is rapid and can be applied in

numerous stages – in coagulation, during washing and stretching or after

stretching. In contrast, dry-spun fibers have an insufficiently porous structure in

the gel state to facilitate producer-dyeing [6].

2.3.3 Pigmented

Among other characteristics, acrylics are lauded for their lightfastness.

This reputation can be maintained without sacrificing strength through pigments.

High acrylics tend to have better light fastness than modacrylics as the latter can

lose halogen groups upon exposure to heat and light. Further, hydrogenated

halogens become brittle and discolor due to induced nitrile polymerization. High

comonomer content also makes the fiber more permeable to oxygen. Oxygen

causes chain scission by reacting with weak bonds on the backbone forming

hydroperoxides.

Pigments are added into the spinning solution as a dispersion in a dilute

polymer solution. Individual suspensions or one suspension can be added. No

significant alteration to the spinning process is required [6].

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 16

2.4 Technical Applications

2.4.1 Reinforcing Fiber (Asbestos Replacement)

Asbestos fibers were well suited for mechanical reinforcement

applications. Documented health hazards spurred interest in identifying

alternatives. Important considerations include tenacity and modulus, retention of

modulus in warm-moist conditions, resistance to alkali degradation and ability of

the fiber to bond to cement. To acquire high tenacity over 70 cN·tex-1, high

molecular weight (100,000 to 500,000) and high orientation are required.

Modifying agents in modacrylics can significantly reduce the mechanical

performance of fibers. Gel spinning acrylic fiber has yielded molecular weights

over 1 million, but cost prohibits any commercial reinforcement applications [6].

2.4.2 Carbon Fibers

Acrylic fibers are a prominent precursor for carbon fiber, over rayon and

pitch, differing by their achievable properties. The fiber must be stabilized in

order to enable its conversion to a cyclic carbon structure. Typically, carboxylic

acids or vinyl bromide copolymers cause ionic initiation and both lower the

temperature for cyclization and moderate the exotherm. A more balanced

exotherm prevents heat buildup and charring in yarn bundles. Effective

monomers include acrylic acid, methacrylic acid, itaconic acid and vinyl bromide.

As in reinforcing fibers, having superior mechanical properties of the fiber

precursor is highly advantageous as it also generates stronger carbon fibers.

Thermoplastic Polyacrylonitrile
Polyacrylonitrile Fiber Technology 17

High molecular weights of over 1 million are sought. High orientation is also

required to enable the chains to contribute completely to the overall strength of

the fiber. All fiber spinning technologies are utilized with tradeoffs between

productivity and fiber denier [6].

2.4.3 Moisture Absorbent Fibers

In many applications, the comfort of a fiber is evaluated on its ability to

absorb moisture without feeling damp or to wick moisture away to an outer

surface where it is evaporated. Dunova® by Bayer was one of the first fibers to

enable this mechanism with a core of thousands of pores and channels with a

protective skin. It absorbs a nontrivial amount of water without feeling wet [6].

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 18

3 Fiber Spinning Technology

Traditionally, high acrylics and modacrylics are solution-spun using either

‘wet’ or ‘dry’ spinning conditions. The first form of solution-spun fibers appeared

in the mid 1850s with the wet spinning of cellulosics into Chardonnet silk. In this

process, after nitrating cellulose with a sulfuric acid/nitric acid solution, it was

dissolved in a solvent to form a viscous ‘dope.’ The gel was forced through a

spinneret forming fibers of cellulose nitrate. Afterwards, the fiber is washed,

stretched and denitrified in an ammonium hydrosulfide bath. This application

illustrates a general method of wet-spinning fibers – convert the fiber to a soluble

form, induce a phase separation to form the fibrous network, and apply an after

treatment to solidify to the desired structure. This process is still used today in

the manufacture of rayon [7].

Two other methods of solution spinning exist: 1) insoluble structures that

are formed after spinning via a reaction of soluble parts and 2) unchanged

soluble chemical forms. Formation of polyvinyl alcohol fibers is an example of

fiber reactant wet spinning. Traditional polyacrylonitrile represents the most

commercially significant (volumetrically) solvent-spun product. This is

accomplished in either a ‘wet’ or ‘dry’ condition. The unifying characteristic of all

solvent spinning systems is that the solvent is miscible in water. This facilitates

its easy removal after spinning [7].

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 19

3.1 Wet Spinning

Wet-spinning systems account for nearly 85% of the worldwide acrylic

fiber production. Many factors contribute to this dominance. A wide array of

solvents can be used, depending on raw material, polymer solubility, after

treatment and economics. The process is noted for its large production capacity

in a relatively safer human environment with low potential for human exposure.

Finally, the fiber in gel form is porous, allowing for in-line dyeing [8].

3.1.1 Process

In the wet-spinning process, dissolved polyacrylonitrile is pumped through

a spinneret that is submerged in a coagulating bath. The bath contains the

solvent and water. As the polymer solution passes through the coagulant or

nonsolvent, a phase change occurs whereby the solvent diffuses out and the

nonsolvent diffuses in. The newly formed fibers emerge in a gel form from the

bath where they are later subjected to a series of aftertreatments.

3.1.2 Solvents

The most commonly used solvents are dimethylformamide (DMF),

dimethylacetamide (DMAc), sodium thiocyanate, and nitric acid, occupying

approximately 24%, 24%, 20%, and 10% of production capacity, respectively [7].

Resulting fiber structures are determined by what solvent, polymer and coagulant

are used. These choices have corresponding options for after-treatment steps.

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 20

3.1.3 Polymer

Solution properties and practical considerations limit the polymer

concentrations. Polymer solubility and solution spinning pressure limits high

concentrations. The ability of a solvent to maintain a spinning solution without

gelation limits polymer solubility. Commercial systems tend to enhance gelation,

so the polymer concentration limit in DMF and DMAc is 30%. High molecular

weight and concentrations create high solution viscosity. Thus, equipment

design limits line speeds and coagulant temperature. However, too low

concentrations prevents proper coagulation and proves too costly for solvent

recovery [7].

3.1.4 Coagulation

Fiber differences from wet and dry spinning result from the stage where

the shape is formed – the coagulation bath. A wet spinneret uses from 20,000 to

100,000 capillaries ranging in diameter from 0.05 to 0.25 mm. Polymer content,

flow rate and temperature are the key parameters that control the product

behavior. Shortly after the emergence of the gelated structure from the

spinneret, physical chemistry mechanisms of heat transfer, mass transfer, fluid

mechanics and solution thermodynamics govern the phase transition as the dope

stream becomes the fiber precursor. In particular, solution and bath temperature

affect the rate of heat transfer, rate of diffusion and phase separation. Polymer

and coagulant concentrations control the rate of diffusion. Diffusion rates for the

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 21

solvent and nonsolvent as well as the phase separation characteristics drive the

underlying transition.

As numerous factors contribute to the ultimate fiber properties, a wide

range of fibers can be formed. Particularly, shape, surface texture, luster and

physical properties can be altered. Even the wet process itself can be changed.

The dry-jet wet spinning procedure moves the spinneret prior to the coagulant,

allowing for a difference in dope and bath temperatures. This procedure also

reduces the fiber stress at the spinneret-coagulant interface of wet spinning.

Dry-jet wet spinning is suited for high production rates with a smaller number of

capillaries. The resulting structure resembles dry-spun fibers [7].

3.1.5 After treatment

Efforts to remove residual solvent and orient the fiber structure ensue

upon emergence of the dope stream from the coagulant bath. Such after

treatments consist of four general stages: washing, orientation, finishing and

relaxing.

Washing is another diffusion-driven stage, controlled by temperature and

concentration of wash liquid. Near-boiling water is typically used as the wash

bath. Generally, this stage occurs either before or in conjunction with stretching.

The fiber is purposefully left with residual solvent and a porous structure to

enable further processing in gel form.

Orientation changes the internal structure of the fiber and yields the final

fiber properties. This occurs at a fiber temperature elevated above the wet glass

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 22

transition temperature where the fiber is stretched 3 to 12 times between rolls

using water as the heat-delivery mechanism.

Finishing is used to facilitate further processing in the final relaxing stage

using lubricants and antistatic chemicals. Composition of the finish is dependent

upon end use. If applied before relaxation, finish chemicals inevitably diffuse into

the structure of the fiber via pores.

The relaxation stage also involves drying and collapsing. The sequence

of these steps depends on the fiber structure generated in the coagulant bath.

Invariably, the fiber collapses due to water removal and the pore structure is

closed. Crimp and relaxation are later applied to give cohesion and bulk to

yarns. Actual relaxation is the final change in fiber morphology through a hot,

wet treatment, which finally sets the mechanical properties of the fiber. Wet

crimp is followed by a one-step dry/collapse/relax treatment. Water is trapped

within the fiber pores as the fiber temperature is driven above the wet glass

transition. The fiber collapses and relaxes by means of water surface tension.

Specific finishing stages are highly cost driven. As the process is dependent

upon fiber structure in the spin bath, characteristics of the

polymer/solvent/nonsolvent system ultimately determine the fiber cost [7].

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 23

3.2 Dry Spinning

In contrast to wet spinning where nonsolvents diffuse into and solvents

diffuse out of the dope, formation of fibers by dry spinning occurs from solvent

evaporation. Evaporation is encouraged through a hot inert gas as the gel

passes out of the spinnerets. I.G. Farben AG first filed patents on the dry-

spinning use of DMF in 1942 as a solvent for polyacrylonitrile, which predated

patents filed by DuPont by only two months. World War II prevented Bayer from

producing it until 1954. DuPont and Bayer’s rights to the dry spinning procedure

created the impetus for other chemical companies to develop the wet spinning

procedure [4].

3.2.1 Process

A properly-mixed and de-aerated solution is pumped through the

spinneret. The most important stage in fiber spinning is the evaporation of the

solvent after the solution emerges from the spinneret. In dry spinning, the

spinneret is ring shaped to circulate the drying gas through the inner mount and

outside. The spinneret is made from stainless steel having generally 2800

apertures.

Once outside the spinneret, the solution quickly turns into gel as it

emerges in the spinning tube or cell. Here, the 6-10 m long, 250-450 mm wide

tube is heated using a jacket of diphyl (26.5% diphenyl, 73.5% diphenyloxide).

The drying gas, usually inert or saturated steam, is heated to 300-350°C flowing

with or against the direction of the gel. Gas must be circulated slowly, 1 – 2

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 24

m·sec-1, to avoid fibers from sticking together. Expected output is 8 – 15kg of

PAN solids per hour with up to 100,000 holes per spinneret and linear speeds of

100 m·min-1. Around 5 – 25% of the solvent is kept in the fiber to aid in further

processing. After sufficient solvent is evaporated, an oil finish is applied to the

filaments, drawn off by godet rollers and wound at 200 – 500 m·min-1 [4].

3.2.2 Solvents

For successful dry spinning, a number of factors must be considered.

First, the polymer must have sufficient stability at the solvent boiling point.

Secondly, the solvent should dissolve well, and not react with dissolved polymer

and have a low boiling point. Additionally, the heat of vaporization must be kept

low while having sufficient thermal resistance, low toxicity, low tendency to

induce static charge, low risk of explosion and easy to recover. Finally, cost

should not be prohibitive. Given these considerations, DMF was considered the

best solvent [4].

3.2.3 Polymer

The solvent content and temperature prevent orientation during the

spinning process. The overall degree of orientation during spinning increases

with the solution viscosity, spinning speed and spinning draft. It can also

decrease with time during material solidification forming microfibers. For product

uniformity, temperature must be well controlled in the spinning solution, spinning

head, drying gas and gas jacket [4].

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 25

3.2.4 Evaporation & Cross Sectional Shape

Solvent evaporates as the solution leaves the spinneret. Once in the hot

atmosphere of the spinning tube, the rate of primary evaporation must be

maintained slowly to prevent a sudden pressure drop. This can cause vapor

bubbles in the filament, which retards the fiber tensile properties.

Secondary evaporation occurs through diffusion. Material, concentration

and temperature-related diffusion coefficients should be controlled throughout

evaporation. Environmental conditions, take-up speed and length of time in

spinning tube determine the amount of solvent present in the gelled fiber.

Differential evaporation and diffusion rates create a solvent gradient with

the outer sheath solidifying faster. Further diffusion from the core reduces its

mass, causing the sheath to collapse inward. Equivalent diffusion and

evaporation rates create a circular cross section. A much faster evaporation rate

creates the bi-lobal, or dog-bone shape. Dwell time in the spinning tube also

affects the cross sectional shape. Longer time in the cell allows for more

diffusion to take place and more prevalent lobes [4].

3.2.5 After treatment

Final fiber properties are imparted in the last stages of processing.

Drawing, washing, finish and drying are the typical steps. Afterwards, fibers are

crimped and packed.

Drawing gives the as-spun fibers the orientation necessary for proper

mechanical properties including tenacity and breaking elongation. Fibers with

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 26

sufficient solvent content do not show a draw point, but must be stretched

adequately to impart full possible orientation. Fibers are stretched in boiling

water between one or two draw zones.

Washing begins before drawing where a tow can be exposed longer to the

washing bath. Washing is conducted under tension. Solvent diffuses out of the

fiber and into the water bath. Circulating the bath keeps solvent concentration

and temperature constant. Final residual content is less than 1% and generally

requires 20 washing zones and a total treatment time of 300 seconds. Finish

lubricates the fibers for further processing, imparts adhesion and prevents static

buildup.

Drying is necessary after washing to remove water. Final water content

should be 1 – 2%. PAN tows must be dried under tension to prevent shrinkage.

Contact heating via heated rollers is the preferred method of drying. Setting

begins at this stage, the degree of which depends on temperature and time and

is minimized to enable crimping. Crimping is generally conducted in a stuffer box

and provides sufficient adhesion between fibers. Tow is fed, heated, and the

fiber is forced to uniformly set [4].

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 27

3.3 Melt Spinning

3.3.1 Conventional Thermoplastics Processing

The three main stages for melt spinning polymers are preparation of the

melt, extrusion and winding. A molten polymer is forced through a spinneret

orifice at a given temperature, pressure and rate. The flow is collected at a

different velocity at the site of ‘take-up.’ The distance between the spinneret and

take-up is variable. Once the polymer reaches the take-up area, the process of

initial fiber formation through solidification and cooling is finished.

From high production rates and absence of required equipment, of all

types of fiber spinning methods, melt spinning is the most economical. The only

caveat is having a thermally-stable and coherent polymer in molten stream form.

Unlike wet- and dry-spinning, there is generally no mass transfer between

polymer and surroundings. Steady state is then defined as the product of

density, cross-section area and velocity of the spinning line at a given distance

from the spinneret. Numerous variables are defined that influence the properties

of the fiber. Ziabicki outlines constituent parameters of three categories: 1)

primary, 2) secondary and 3) ‘resulting’ [9].

Primary variables are concerned with predetermined technical settings

that establish the foundation for the extrusion. The most important primary

variable is the material thumbprint such as chemical composition, molecular

structure and physical behavior. Other associated variables are the extrusion

temperature, dimensions and number of spinneret orifices, mass outflow, length

of spinning path, take-up velocity and cooling conditions.

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 28

The secondary variables relate to the equation of continuity between

density, area and velocity. In particular, extrusion velocity, average diameter and

titre of the as-spun filament and the spin-draw ratio, or ratio of take-up velocity

and extrusion velocity, are the defined variables.

The resulting variables follow from the kinematics and dynamics of the

spinning process. These characteristics include take-up tension, take-up stress,

ultimate filament temperature, and any other texture or physical properties of the

as-spun fibers [9].

3.3.2 Imparted melt processability

For PAN, melt processability can be imparted if the melting point is

brought sufficiently below the degradation point. One of the most direct ways is

inclusion of 5 – 20% of low molecular weight polyacrylonitrile or incompatible

lubricating component such as a polyolefin or polyamide. However, these

materials can adversely affect the mechanical properties and chemical

resistances of the resulting fiber.

Copolymerization to reduce the melting point has been investigated since

the early 1950s. However, the distribution of comonomer in the copolymer

structure allowed in traditional free radical polymerization reactions prevented

success. One successful attempt utilized acrylic esters or methacrylic esters in

emulsion polymerization, where after fiber conversion, 11% of the polymer was

extracted with acetone. The melt was spun at 1500 m·min-1 at 200 ºC with a 1.8

denier fiber and tenacity of 4.5 gf·den-1.

Thermoplastic Polyacrylonitrile
Fiber Spinning Technology 29

3.3.3 PAN Hydrates

Water decouples the dipole interaction between the acrylonitrile

sidegroups and has been employed in an attempt to produce melt-spun fibers

from a high pressure (30 to 70 bar) system. PAN exists in four primary phases,

depending on temperature and contact with water: solid, hydrate, dual and

suspension. At temperatures above 150ºC and water contents between 5 and

25%, the hydrate form is a metastable single phase where water evaporates out

of the hydrate stream to form fibers. The throughput rate is thus a function of the

rate of evaporation.

Mechanically, the fibers have a tenacity of about 4.5 gf·den-1 and a

breaking elongation of 40%. This fiber also has an irregular cross-section.

Quality as well as economic issues, such as pressure, throughput and energy

consumption, are obstacles that prevented implementation of this technology [4].

3.3.4 Summary of Prior Art in ‘Melt Processable’ PANs

Over recent decades, foreign and U.S. patent offices granted hundreds of

patents relating to melt processability of high acrylonitrile-content polyacrylonitrile

and related variants. Nearly all, except for those relating to Amlon®, focus on

imparting melt processability through copolymerization with post-conversion

solvent extraction, reduction of melt viscosity by inclusion of low-molecular

weight acrylonitrile polymers, copolymerization to reduce melting point, PAN

hydrates, and inclusion of incompatible lubricants and other plasticizers.

Thermoplastic Polyacrylonitrile
Markets and Potential 30

4 Markets and Potential

A thermoplastic version of high polyacrylonitrile has been sought for many

decades. However, the anomalous paracrystalline morphology and lack of ability

to control the structure of the copolymer prevented an innately thermoplastic form

from being available. Now that a process for structuring the polymers has

emerged, the larger potential for the material can be explored.

4.1 Trends in the Worldwide Synthetic Fiber Production

Worldwide consumption of synthetic fibers shows clear trends for the past

10-15 years, Figure 2. Polyester, or poly(ethylene terephthalate), continues to be

the largest fiber in production and the fastest growing. Filament rather than

staple is the most popular form. Polyolefins including poly(ethylene) and

poly(propylene) show steady growth since 1994. Aramids, nylon (PA) filament,

acrylic, and cotton fibers show nearly constant consumption.

4.2 Trends in the Acrylic Fiber Market

In the decades following the introduction of acrylic fibers by DuPont in the

1950s, the U.S. marketplace grew rapidly. Only years after the release of

Orlon®, six other companies released their own acrylic fiber: Union Carbide’s

Dynel®, Monsanto’s Acrilan®, American Cyanamid’s Creslan®, Dow Chemical’s

Zefran®, Tennessee Eastman’s Verel® and Goodrich’s Darvan® [10]. The primary

uses of acrylic fibers were for wool replacements in carpet, bulk fabrics and

blankets.

Thermoplastic Polyacrylonitrile
Markets and Potential 31

25

Production (millions of metric tons)


20

15

10

1994 1996 1998 2000 2002


Acrylics Year Nylons
Polyester Polypropylene

25
Production (millions of metric tons)

20

15

10

1985 1990 1995 2000 2005


Acrylic & Modacrylic Nylon & Aramid
Year
Polyester Polyolefin

Figure 2. Synthetic Fibers Market: Acrylics, Polyester,


Nylons, Aramids, and Polyolefins

a) Worldwide Consumption [11]


b) Worldwide Consumption, [12-19], a excluding U.S. Aramids

Thermoplastic Polyacrylonitrile
Markets and Potential 32

4.2.1 Markets in the United States

After the initial peak in usage, the market subsequently declined in the

early 1970s as nylon outperformed acrylic in terms of elastic recovery. Given the

type of wear in carpet applications, recovery of the fiber to its original shape is

highly desirable. The first large overcapacity for acrylic fiber production ensued.

Consequently, producers shifted their efforts to the apparel fibers market, which

itself peaked in the 1980s. However, the popularity of cotton in sweaters and

socks caused the second decline in the acrylic fiber market. Due to a gross

world overcapacity, DuPont began closing foreign manufacturing sites in 1978,

which continued until 1990 when it exited the acrylic business altogether.

Producers became increasingly dependent upon foreign exports to Western

European and Japanese markets, Figure 3. However, consumption in foreign

markets gradually declined, Figure 4, and domestic operating capacity followed

suit. Gross worldwide overcapacity eroded profit margins and producers of

commodity acrylic fibers closed operations, spun-off companies or sold assets.

During the late 1990s, the remaining descendants of the first producers of acrylic

fibers – American Cyanamid and Monsanto – divested their fiber businesses.

Sterling Chemicals [20], Inc. purchased Cytec’s assets in 1997, which itself was

a 1993 spin-off of American Cyanamid [21]. Solutia, Inc., was Monsanto’s 1997

chemical spin-off [22] and until 2005, they were the only two major commodity

acrylic fiber U.S. producers [23]. On February 2, 2005, Solutia closed its

remaining production site, exiting the acrylic fiber business [24].

Thermoplastic Polyacrylonitrile
Markets and Potential 33

Amount (thousands of metric tons)


400

300

200

100

1975 1980 1985 1990 1995 2000 2005


Year

Capacity Production Domestic Shipments


Imports Exports Consumption

Figure 3. Capacity, Production, Domestic Shipments, Imports, Exports and


Consumption in United States Acrylic Fiber Market.
[23]

Thermoplastic Polyacrylonitrile
Markets and Potential 34

600

Amount (thousands of metric tons)


500

400

300

200

100

1980 1985 1990 1995 2000 2005


Year

1500
Amount (thousands of metric tons)

1250

1000

750

500

250

1980 1985 1990 1995 2000 2005


Year
Capacity Production Consumption

Figure 4. Capacity, Production and Consumption in the Acrylic Fiber


Markets by Major Geographical Regions

a) Asia [23].
b) China [23].

Thermoplastic Polyacrylonitrile
Markets and Potential 35

600

Amount (thousands of metric tons)


500

400

300

200

100

1975 1980 1985 1990 1995 2000 2005


Year

500
Amount (thousdands of metric tons)

400

300

200

100

1970 1975 1980 1985 1990 1995 2000 2005


Year
Capacity Production Consumption

Figure 4. Capacity, Production and Consumption in the Acrylic Fiber


Markets by Major Geographical Regions. (continued)

c) United States [23].


d) Japan [23].

Thermoplastic Polyacrylonitrile
Markets and Potential 36

1250

Amount (thousands of metric tons)


1000

750

500

250

1970 1975 1980 1985 1990 1995 2000 2005


Year
Figure 4. Capacity, Production and Consumption in the Acrylic Fiber
Markets by Major Geographical Regions (continued)
e) Western Europe [23].

250
Amount (thousands of metric tons)

200

150

100

50

1975 1980 1985 1990 1995 2000 2005


Year

Capacity Production Consumption

Figure 5. Acrylic Fiber Consumption by Apparel, Home


Furnishings and Industrial Market Segments.
a) United States [23].

Thermoplastic Polyacrylonitrile
Markets and Potential 37

100

Amount (thousdands of metric tons)


75

50

25

1990 1995 2000 2005


Year

500
Amount (thousands of metric tons)

400

300

200

100

1985 1990 1995 2000 2005


Year

Apparel Home Furnishings Industrial & Other

Figure 5. Acrylic Fiber Consumption by Apparel, Home


Furnishings and Industrial Market Segments. (continued)

b) Japan [23].
c) Western Europe [23].

Thermoplastic Polyacrylonitrile
Markets and Potential 38

25

Production (millions of metric tons)


20

15

10

1994 1996 1998 2000 2002


Acrylics Year Nylons
Polyester Polypropylene

25
Production (millions of metric tons)

20

15

10

1985 1990 1995 2000 2005


Acrylic & Modacrylic Nylon & Aramid
Year
Polyester Polyolefin

Figure 6. Synthetic Fibers Market: Acrylics, Polyester,


Nylons, Aramids, and Polyolefins

a) Worldwide Consumption [11]


b) Worldwide Consumption, [12-19], a excluding U.S. Aramids

Thermoplastic Polyacrylonitrile
Markets and Potential 39

4.2.2 Markets in Europe and Asia

Consumption of acrylic fibers in the traditional markets – the United

States, Western Europe and Japan – continues to contract. Strong exports

characterize these markets, where traded fibers are more application specific.

The use of acrylic fibers in Western Europe tends to be bulk fabrics, Figure 5.

Japanese use is mainly carbon fibers. The balance of these exports finds buyers

in China, where consumption is rapidly accelerating. In 2000, China became not

only the largest consumer of acrylic fibers worldwide, but consumed more acrylic

fibers than all other combined acrylic worldwide markets with 840,000 million

metric tons. While their production is increases, the gap between consumption

and production grows. While the demand to import fibers into China is

increasing, the pervading worldwide overcapacity prevents some from remaining

in operation. Indonesia shows the second largest consumption of acrylic fibers at

120,000 metric tons in 2004 [23]. India is the third largest Asian market, whose

annual consumption of acrylic fibers was only 102,000 metric tons, or

approximately 8.5% of China [25].

4.2.3 Market Prospects

In a market with this history, a new acrylic fiber must have substantial

benefits. Cost savings will be an important aspect, as solution processing

contains inherently more costly procedures than melt processability. Other

factors for success in fiber markets are improved mechanical properties. In

product markets where acrylics are widely used, such as outdoor fabrics,

Thermoplastic Polyacrylonitrile
Markets and Potential 40

manufacturing textiles from filaments rather than staple yarns should decrease

the overall production costs. If the fiber performs better, specifically in tenacity,

washability, crease-ability, and elastic recovery, low raw materials cost could

make it a less-expensive alternative to polyesters and/or nylons. Furthermore, in

addition to fibers, a melt-processable material will offer many new alternatives in

the wider thermoplastic polymer markets of films. Polyacrylonitrile would then

offer its unique properties – namely UV and chemical resistances – and improved

mechanical performance in a commodity, thermoplastic form.

4.3 Cost Analyses

The benefits of melt processing, aside from mechanical enhancements,

are cost driven. The least expensive manufacturing route is synthesis of the

polymer followed by direct conversion to fibers. Cost predictions of melt

processable high acrylonitrile are based on raw materials and estimated

conversion costs for poly(ethylene terephthalate), polypropylene, and polyamides

or nylons. The results are presented in Table 1.

4.3.1 Poly(ethylene terephthalate)

Poly(ethylene terephthalate) is synthesized from ethylene glycol (EG) and

dimethyl terephthalate (DMT) or purified terephthalic acid (PTA). Polymerization

with PTA is now the dominate method because of shorter reaction times and

reduced amounts of required catalyst. Polyester fibers have the largest

worldwide demand with approximately 20.9 million metric tons in 2002.

Thermoplastic Polyacrylonitrile
Markets and Potential 41

Approximately 0.39 pounds of EG and 0.90 pounds of PTA are required per

pound of synthesized PET [26, 27]. Raw material, resin, filament and staple

costs are shown in Figure 7 a). From 2000 – 2003, the average conversion cost,

or difference between filament and raw material costs, was 27 cents·pound-1.

4.3.2 Poly(propylene)

Polypropylene, a polyolefin, is synthesized directly from propylene with

assistance of metallocene catalysts that enable molecular weights and bulk

properties to be engineered. It is one of the easiest polymers to handle and

fastest growing segments of the synthetic fiber industry with 2002 production just

under 6 million metric tons [28, 29]. Approximately 1.01 pounds of polymer-

grade propylene are required for each pound of polypropylene. Raw material,

resin and various fiber and yarn costs are shown in Figure 7, b). Since 2000 the

average conversion cost, or difference between raw materials and polymer resin

was 15 cents·pound-1. A lower conversion cost, when compared with PET, is

due to the difference in final product (resin versus filament) and the higher

operating temperature required to process molten PET.

4.3.3 Poly(amides)

Polyamides, or nylons, are a general term for related chemical structures.

They are named for the number of carbon atoms between the two amide bonds.

Nylon 6 and Nylon 66 are the predominate forms with somewhat different

properties, accounting for nearly 98% of the worldwide nylon production. Nylon 6

Thermoplastic Polyacrylonitrile
Markets and Potential 42

is synthesized directly from caprolactam requiring 1.01 pounds per pound of

Nylon 6. Nylon 66 is synthesized from adipic acid and hexamethylenediamine

requiring 0.73 and 0.57 pounds per pound of polymer, respectively [30]. Nylon

fibers have some of the highest tenacities of commodity fiber-forming polymers,

ranging from 4.5 – 5.5 gf·den-1. In 2000, over 4 million metric tons of nylon fiber

was produced. Raw material, resin, and fiber costs are shown in Figure 7, c).

4.3.4 Solution processable high acrylonitrile copolymers

While the cost of raw polyacrylonitrile materials – chemical grade

propylene and ammonia – is the lowest raw material cost among the major

thermoplastic fiber-forming polymers. Trends in raw materials, intermediate and

fiber costs are shown in Figure 7, d). In 2002 the cost was approximately 19

cents·pound-1 [31], versus 20 for the polymer-grade propylene used in

polypropylene. However, 1.2 pounds of chemical grade propylene are required

for one pound of acrylonitrile, bringing the raw material cost to 22 cents·pound-1

versus a similar 20 cents·pound-1 for polypropylene, whose difference changes

where the cost of chemical and polymer grade propylene vary somewhat

independently.

Solution-spun fibers tend are relatively expensive, and acrylic fibers cost

approximately 85 cents·pound-1 in 2002 [23]. Low throughput ( 200-500 m·min-1

[4]) and solvent recovery are the primary factors that drive the higher prices.

Since 2000, the cost of conversion from raw materials to fiber was 58

cents·pound-1.

Thermoplastic Polyacrylonitrile
Markets and Potential 43

4.3.5 Melt processable high acrylonitrile copolymers

In acrylic copolymers, inclusion of comonomers will increase the cost of

the resin. In 2003, the cost of vinyl acetate was 48 cents·pound-1 [32] and methyl

methacrylate 65 cents·pound-1 [33]. In 1994, methyl acrylate was 12

cents·pound-1 more expensive than methyl methacrylate [34], putting the current

price probably between 70 to 80 cents·pound-1. While methyl acrylate and vinyl

acetate are similar chemicals, despite its higher cost, methyl acrylate was the

original comonomer used in Orlon® [10], is still used in Dralon® and is the

comonomer used for the Amlon® D copolymer in this study. Based on the most

recent prices, the raw material cost for an 85% acrylonitrile acrylic copolymer

would lie between 29 and 31 cents·pound-1. In 2002, the raw material costs of

poly(ethylene terephthalate) was 36, nylon 6 72 cents·pound-1, and nylon 66 117

cents·pound-1. Thus, the raw materials for acrylonitrile are the second least

expensive of commodity synthetic fibers.

While a more detailed costing of acrylic fibers by comonomer and

acrylonitrile content was not given, this conversion cost should be considered an

estimate for solution processing acrylics. The conversion cost for the other

synthetic fibers depends on the continuity of production. Since 2000, the mean

conversion cost or difference in fiber and raw material costs for solution-

processed acrylic is 58 cents·pound-1. Likewise, the mean conversion cost of

raw materials to resin was 40 cents·pound-1 for unsaturated poly (ethylene

terephthalate), 15 cents·pound-1 for polypropylene, nylon 6 53 cents·pound-1,

nylon 66 21 cents·pound-1, and polypropylene 15 cents·pound-1. Given these

Thermoplastic Polyacrylonitrile
Markets and Potential 44

costs of converting the raw materials to a resin and the raw material cost for

acrylonitrile, the expected cost of melt processable polyacrylonitrile resin might

be between 45 and 65 cents·pound-1, depending on the efficiency of

polymerization and type of comonomer produced.

Table 1. Raw Materials and Conversion Costs for Major Thermoplastic Polymers
Raw Cost Conversion Costs (¢·lb-1)
Resin Materials
(¢·lb-1) Resin Fiber
Ethylene glycol and
Stp: 19
PET (dimethyl terephthalate or 36 40
Fil: 27
purified terephthalic acid)
PP Polymer grade propylene 22 15 n/a
6: Caprolactam
6: 72 6: 53
PAs 6,6: adipic acid and Fil: 21
6,6: 117 6,6: 21
hexamethylenediamine
Chemical grade propylene
coPNAs ±30 n/a Stp: 58
and ammonia
Chemical grade propylene Stp: 20 – 30
Amlon® ±30 45 - 65
and ammonia Fil: 30 – 45

For carpets, where polypropylene and nylons dominate, the cost of

conversion is approximately 78 and 74 cents·pound-1. Thus melt processable

polyacrylonitrile carpet yarns are expected to cost between 110 and 125

cents·pound-1, putting it between the low end and high end markets where

properties and cost may expropriate market share the competitors. Since 2000

for fibers, the mean difference in filament and raw materials cost for

poly(ethylene terephthalate) was 27 and staple 19 cents·pound-1. The difference

in an undesignated type of nylon in staple form and nylon 6 raw material cost

was 21 cents·pound-1. Given these costs and based on the estimated raw

Thermoplastic Polyacrylonitrile
Markets and Potential 45

material cost for acrylonitrile copolymers, the projected cost for melt processable

polyacrylonitrile filaments is 60 to 75 cents·pound-1 and staple 50 to 65

cents·pound-1. The staple cost corresponds with the extra cost of solution

processing – approximately 20 – 30 cents·pound-1.

Thermoplastic Polyacrylonitrile
Markets and Potential 46

120

100

Cost (cents per pound)


80

60

40

20

1980 1985 1990 1995 2000 2005


Year
Raw Materials Unsaturated Resin
Filament (150 denier) Staple

250

200
Cost (cents per pound)

150

100

50

1980 1985 1990 1995 2000 2005


Year
Raw Materials Homopolymer Resin
Carpet Face Yarn (2750 denier) Upolstry (fine denier)
Sanitary Applications (1.5-2.2 den) Needlepunched Geotextiles (5-8 den)

Figure 7. Raw Material, Intermediary and Fiber Costs for Poly(ethylene


terephthalate), Poly(propylene), Poly(amides) and Polyacrylonitriles.

a) Poly(ethylene terephthalate) [26, 27, 35, 36].


b) Poly(propylene) [28, 29, 37].

Thermoplastic Polyacrylonitrile
Markets and Potential 47

300

250

Cost (cents per pound)


200

150

100

50

1980 1985 1990 1995 2000 2005


Year
Nylon 6 Raw Materials Nylon 66 Raw Materials Nylon 6 Resin
Nylon 66 Resin Carpet Yarn (1200 den) Textile Yarn (40 den)

125

100
Cost (cent per pound)

75

50

25

1980 1985 1990 1995 2000 2005


Year
Raw Material Acrylonitrile List
Acrylonitrile Contract Acrylic Fiber (3 denier)

Figure 7. Raw Material, Intermediary and Fiber Costs. (continued)

c) Poly(amide), Nylons [30, 38-41].


d) Poly(acrylonitrile) [23, 29, 31].

Thermoplastic Polyacrylonitrile
Markets and Potential 48

4.4 Potential Applications for Thermoplastic PAN

Historically, high cost, low throughput and difficult processing prevented

polyacrylonitrile from reaching its full potential as a commodity polymer. Few

unmodified vinyl polymers offer such characteristic properties, especially those of

chemical and UV resistances. Fluorinated and chlorinated chains are notable

alternatives, each having their specific uses, which are dependent upon their

unique chemistries.

4.4.1 Comparative properties between synthetic fibers

Poly(ethylene terephthalate) filament and staple fibers are used as the

benchmark, because they dominate the synthetic market and its tenacity and

density are considered relatively high, Table 2 and Figure 2. It has good UV

resistance, abrasion resistance and strength, and exceptional wash-wear

performance, wrinkle resistance and resistance to burning [42]. Nylons are

specialty fibers finding use in industrial applications and products where elastic

recovery is important. They are known for exceptional abrasion resistance,

strength and elastic recovery, but poor resistance to UV radiation. Polypropylene

is an inexpensive, easy-to-handle fiber with standard mechanical properties.

They are difficult to dye and are unable to hold creases. Acrylics are not

traditionally known for high strength. However, they do exhibit excellent hand,

bulk, wick-ability, UV and chemical resistances.

Thermoplastic Polyacrylonitrile
Markets and Potential 49

4.4.2 Deniers, profiles and multicomponent fibers

Properties of fibers and fabrics depend not only on the chemistry of the

fiber, but often times more importantly on their physical structures. A large

diameter, for instance yields a higher bending rigidity, which is helpful in some

mechanical applications. However, fine denier fibers are useful for good draping

and soft hand. Finally, optical properties are dependent not only on the diameter,

but on the shape of the fiber. The latter affects the way light is reflected or

refracted, which creates optical illusions.

Table 2. Comparative Properties of Synthetic Fibers


Tenacity Elongation Specific Moisture
Fiber
( gf·den-1 ) (%) Gravity Regain (%)
Poly(amide) [30] 4.5 – 5.6 25-33 1.14 4.0 – 4.5
Poly(ethylene terephthalate) [35] 4.6 – 5.0, 8.9 20-55, 12 1.38 0.4
Poly(propylene) [43] 3.0 – 4.0 80 – 100 0.91 0.0
Poly(acrylonitrile) [23] 2.5 – 4.0 20 – 55 1.16 – 1.18 1.0 – 2.5

Traditionally, acrylic fibers are either bi-lobal or round. With a

thermoplastic form, diameters can be shaped as tri-lobal, or any other kind of

profile. Fibers with desired luster can thus be obtained, specifically for carpets.

Already, some Amlon® fibers have been observed to have high lustrous

properties. Since the density of acrylic copolymers, which is dependent upon the

type and concentration of comonomers, is similar to nylon [44], a standard in

fashion fabrics, lightweight lustrous fabrics from Amlon® can also be constructed.

As denier not only affects the mechanical and optimal properties,

decreasing diameter also provides more surface area. Nano- and microdenier

fibers ( < 1 denier per fiber [45] ) are typically made from bi- and multicomponent

Thermoplastic Polyacrylonitrile
Markets and Potential 50

systems. The fibers and fabrics are often used to achieve suede hand, filtration

and in absorptive applications [46]. Amlon® would potentially enable a range of

new technical fabrics from material containing a high content of acrylonitrile with

the characteristic resistant properties.

4.4.3 Reinforcing fiber

Acrylic fiber tow finds application as reinforcement for cement and

concrete products. The polymer should be 100% acrylonitrile to hinder

decreases in tenacity and modulus from warm-moist effects [6]. Meltability of a

fiber may not be advantageous in applications that demand consistent

mechanical behavior in hot and wet environments. Furthermore, high molecular

weights, whose range is between 100,000 and 500,000, are typically used. This

can significantly increase solution viscosity and would likewise increase melt

viscosity. A melt-processable PAN might not be suited in this application.

4.4.4 Carbon fibers

Similar to reinforcing fibers, very high molecular weights of

polyacrylonitrile (and copolymers with low concentrations of comonomers) are

necessary to ensure suitable tenacity and modulus of carbon fibers [6, 47], which

range from 2550 to 4000 MPa and 230 to 600 GPa, respectively. Furthermore,

low content of comonomer (2 – 5%) are in carbon fiber precursors. As such,

viscosity and processability in a thermoplastic form may be problematic.

Thermoplastic Polyacrylonitrile
Markets and Potential 51

Furthermore, the conversion of precursor fibers to carbon fibers involves a

thermomechanical stabilization treatment in order for successful pyrolysis to

carbon. During this stage, the fiber is placed under low tension and slowly

heated to 260 – 290ºC while the polymer chains cyclize to form an imine

structure and then oxidize to form a pyridone structure. In other words, the

sidegroups collapse and cross link to form an intermediate conjugated structure.

The nitrogen in the acrylonitrile sidegroup loses its triple bond to form a double

bond with its partner carbon and forms a single bond with the carbon in the

sidegroup of the neighbor. The cyclized structure is oxidized to maintain its

stability during pyrolysis.

This process of thermomechanical stabilization has emerged specifically

for the traditional solution-processed acrylic fiber. Therefore, it is possible that a

similar mechanism of stabilization may be unsuccessful in converting a melt

processable fiber at temperatures above its melting point. For polyacrylonitrile,

alternative forms of oxidation stabilization involve UV radiation, gamma

irradiation, and chemical treatment, the former was investigated by Paiva for

Amlon® [48].

Pitch is an alternative carbon precursor material made from petroleum and

coal byproducts and finds application in high modulus carbon fibers (100-1000

GPa). As it is melt processed, albeit at a higher temperature, similar stabilization

difficulties are present. Pitch fibers are successfully thermomechanically

oxidized materials, whereby the mechanical properties are best when oxidation is

optimized. For Amlon®, a new procedure for stabilization and oxidation must be

Thermoplastic Polyacrylonitrile
Markets and Potential 52

developed. Guidelines for an appropriate technique may be found in existing

technologies.

As the Amlon® technology includes olefinically unsaturated monomers,

comonomers like itaconic acid that are more appropriate for carbonization may

be investigated. The amount of comonomer and the associated bulk properties

require engineering for specific applications. In summary, the use of Amlon® for

carbon fiber technologies will likely require significant and long-term

development.

4.4.5 Composite and carbon carbon composite components

Amlon® will likely find application in the larger field of composites and

carbon carbon composites. Various thermoplastic, thermoset and epoxy

matrices are employed in a wide range of composite applications from pultruded

to laminated products [49]. The existence of a thermoplastic polyacrylonitrile

would provide a low cost carrier where UV and chemical resistance is

advantageous. In particular, marine, aerospace, automobile, and consumer

product industries are probable candidates.

While Amlon® may require a more specific method of conversion to carbon

fiber, it may also serve as a carbonaceous matrix in carbon carbon composites

(CCC). CCCs are constructed using carbon fibers suspended in a carbon

precursor matrix. Aside from superior mechanical properties in a lightweight

material, CCCs are also known for their thermal and electrical conductivities. As

such, they are employed in various aerospace and high tech applications.

Thermoplastic Polyacrylonitrile
Markets and Potential 53

CCCs are fabricated using two main techniques: gas phase infiltration

(GPI) and liquid phase infiltration (LPI) [50]. In GPI, the composite is heated in a

furnace up to 1100 ºC where hydrocarbon gas is circulated. A “crust” is gradually

deposited in the substrate surface. In LPI, matrices are typically either pitch or a

thermoset resin like phenol formaldehyde. The composite is oxidized and

converted to carbon and densified by filling voids that are created during

degassing of carbonization byproducts within the matrix. Densification usually

involves two or three stages to ensure a consistent structure. Low viscosity

phenol is used in monolithic composites, whereas pitch is impregnated in thin-

section carbonaceous preforms. The material choice depends on the application

and cost. Amlon®, similar to pitch, may offer a matrix alternative for partially or

fully carbonized composite materials where the achievable conductive,

mechanical and resistant properties may be optimized.

4.4.6 Limitations from the hot-wet conundrum

One of the main reason that traditional acrylic failed to maintain a

dominant presence in the apparel and carpet markets is that the fibers are

unable to retain their moduli in hot-wet conditions [5]. This leads to a loss of bulk

and shape after repeated washing and drying. Under humid conditions, fibers

lose their resilience and wrinkles develop that are difficult to remove. Over time,

this causes the fabrics to lose their original shape. This property is attributed to

changes in glass and melting temperatures in the presence of water. Numerous

methods of improving this property have been attempted with limited success.

Thermoplastic Polyacrylonitrile
Markets and Potential 54

4.4.7 Outdoor fabrics

Due to the exceptional UV resistance in high acrylic fibers, their fabrics

find use in many outdoor applications. Primarily, products include awnings,

convertible automobile covers and marine textiles woven mainly from ring-spun

yarns. A melt-processable acrylic would enable these products to be made less

expensively through lower material costs, use of filaments rather than staple

fibers, or through nonwovens fabrics.

4.4.8 Nonwovens

To date, almost all acrylic fiber nonwovens are constructed from wet-lay,

air-lay or carding processes. However, the amount of melt spun nonwoven

production (at least in the U.S. and Canada) has steadily increased [51]. A

thermoplastic polyacrylonitrile would potentially open the melt spun industry to

acrylic for fabrics, filtration [46] and high tech products including carbonized

fabrics. The specific technologies used are dependent upon the product, which

relates to the fiber denier [52]. Cold drawing, a phenomenon required for melt

blown technology, has been observed in Amlon®.

4.4.9 Membranes, Films and Conductivities

Membranes are constructed in many forms for specific properties and

applications [53]. One of the methods that pertains to fibers is hollow fiber

membranes (HFM). HFMs are advantageous because they allow a high

modulus composite with high surface area. The diameter of the fiber ranges

Thermoplastic Polyacrylonitrile
Markets and Potential 55

from 50 to 3000 µm and is typically used in pressurized gas filtration,

hemodialysis or separation of particulate matter. While these fibers are typically

solution-spun, Amlon® may provide an easier method of fabrication.

The application of selectively permeable membranes in filtration for media

and for protection against biological and chemical warfare agents is fast

increasing. For many applications, chemical resistance is a desired property.

The cost and thermomechanical properties of the polymer often restrict the

application. A commodity thermoplastic, polyacrylonitrile may offer a higher

performing, material in terms of chemical and UV resistances, for less cost.

The enhancement of properties through grafting, doping and inclusion of

specific comonomers for electronic conductivity has been demonstrated for

acrylic fibers [5]. In several Japanese patents, electrically conducting fibers have

been made using metal dopants. Furthermore, the inclusion of sulfur

comonomers into acrylonitrile copolymers for improved dye-ability is well known

[34]. Fluorosulfonated semi-permeable membranes known as Nafion® have

dominated the ion exchange membrane market for polymer electrolyte fuel cells

[54]. The material is able to resist the corrosive environment, but its performance

is limited because of the inherent hydrophobicity of the polymer, which reduces

its ability of allowing hydrogen ions to cross the membrane. As polyacrylnotrile is

chemically resistant, while being hydrophilic, it may offer a better performing

alternative.

Regardless of whether the membrane, essentially a thin film, is used to

restrict the passage of chemicals, the gamut of less sophisticated film and

Thermoplastic Polyacrylonitrile
Markets and Potential 56

bottling applications are also possible. Today, materials with less than 85%

acrylonitrile are used in a well known barrier film under the trade name Barex®.

With the resistances unique to high acrylonitrile copolymers, low material cost

and mechanical property enhancements inherent to melt-processability, better

performing filaments, new thermoplastic applications and high tech markets are

likely.

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 57

5 Polymer Microstructure

Polyacrylonitrile (PAN) is a unique vinyl polymer because of its chemical

resistance and semi-crystalline morphology. One of its most notable properties

is its relatively high ‘melting’ point of ~320º C and low thermal degradation range

from 250 - 310º C [55]. For most thermoplastic polymers the relationship

between crystalline morphology and melting point is well understood in terms of

discrete crystalline and amorphous domains and as an equilibrium first-order

thermodynamic phase transition [56]. However, PAN does not exhibit the

classical two-phase morphology and has been described as “amorphous with a

high degree of lateral bonding” [57], or as a “two-dimensional liquid crystalline-

like structure with many defects” [58]. As a result, and despite conflicting

conclusions about its tacticity as radically-polymerized [59-61], the degree of

crystallinity for the homopolymer does not appear to be a function of

stereoregularity [62]. Similar behavior is also observed in poly (vinyl alcohol) and

poly (norbornene) [63].

Unlike other vinyl polymers, where stereoregularity is required to

crystallize [64], this mesomorphic crystalline polymer exhibits lamellar single-

crystal-like platelets [65] and a variety of lattice structure morphologies that are

independent of stereoregularity [55]. It does not exhibit regular ‘melting’

thermograms, except when using differential scanning calorimetry at a high ramp

heating rate of 160 ºC·min-1, where it shows what appears to be an endotherm

prior to exothermic degradation reactions [5].

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 58

The first order thermodynamic melting transformation, is defined by

∆H m
Tm = ,
∆S m

where Tm, ∆Hm and ∆Sm are the temperature, enthalpy and entropy of melting,

respectively. In terms of equilibrium melting, the interchain interactions between

polar nitrile substituents, which must be incorporated in the PAN crystals, are

expected to contribute primarily to the enthalpy of melting. The change in

entropy upon melting, on the other hand, is associated mainly with disordering

the intrachain PAN conformations [66]. Measurements of melting point

depression in dilute solution show that the heat of fusion of PAN is less than that

of poly (chlorotrifluroethylene) and much less than that of polyethylene [67],

implying that substantial interchain interactions between nitrile substituents

persist in the PAN melt. Because of its high Tm and relatively small ∆Hm, the

change in entropy upon melting PAN must be very small [68], which is consistent

with substantial conformational freedom or diversity for the PAN chains even

when incorporated in crystals.

In general, to alter the crystal morphology and melting temperatures of

homopolymers, comonomers are used to reduce the copolymer melting

temperature. Theoretical works by Flory and by Eby suggest microstructural

control of copolymer melting temperatures. Flory asserts that comonomer

interruption and shortening of the length of the crystalline monomer sequences,

reduces the number and average size of crystallites [69]. Because chain

segments on the crystallite surfaces have larger free energy, smaller crystals

with a larger surface to volume ratio have a reduced melting point [58].

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 59

Specifically, the crystallizable monomer sequence length decreases linearly with

the addition of the comonomer:

1 1 R
− o= XB,
Tm Tm ∆H u

where Tm and Tmº are the melting points of the copolymer and homopolymer,

respectively, R is the universal gas constant, ∆Hu is the heat of fusion per mole of

crystalline repeat unit, and XB is the mole fraction of the minor non-crystallizing

comonomer. Furthermore, for non-random copolymers, Flory proposes that the

melting point depression depends on the sequence propagation probability and

not the overall comonomer composition (XB) [70]. Eby, on the other hand,

proposes that the comonomer may be partially incorporated into the crystal

lattice as defects and extends Flory’s theory by adding a parameter that accounts

for the degree of lattice disruption [71].

For PAN copolymers there exists evidence in support of both theories. In

support of the Flory theory, Slade, et al. experimentally verified the relationship

between the depression of melting temperature and comonomer concentration

for vinyl acetate/AN copolymers [72]. Kulshreshtha, et al. showed that

crystallinity and crystal thickness decrease with larger concentrations of methyl

acrylate comonomer [73], until the structure is highly disordered [74]. Frushor

demonstrated evidence in support of the Eby theory with a melting point

depression constant, which indicates the degree to which comonomers disrupt

the crystalline lattice, based on the molar volume of their sidegroups [74].

To date, the only reported way to affect the PAN melt temperature is

through controlling the parameters of free radical polymerization, including

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 60

monomer concentrations and reactivities. The enchainment of comonomers and

their sidegroups in the resulting copolymers are considered to be sequenced in a

somewhat blocky manner, and are not completely randomly inserted [75]. At a

comonomer concentration over 15%, methyl acrylate or vinyl acetate create a

modified acrylic, or modacrylic, which at sufficiently high incorporation produces

an amorphous glassy material. To the extent that the thermal properties are

dependent upon the length of the acrylonitrile sequences, a melt processable

high PAN (high AN content) is possible if the lengths of the AN sequence were

controllable, such that they are long enough to impart crystallinity, while keeping

the melting point below the degradation range. This would support the proposal

by Flory that in this non-random copolymer, the sequence propagation

probability, which determines the average AN length, controls the melt behavior.

The crystalline domains could, however, still include the comonomer as indicated

by Eby and Frushour.

British Petroleum (BP) scientists Smierciak [76], et al. developed a novel

‘starved’ or ‘scavanged’ emulsion copolymerization reaction based on work by

Dimitratos [77]. Under extremely low feed rates, the polymerization rate

approaches a pseudo-steady value based on the availability of the monomer and

its addition rate, and the composition of the copolymer is similar to the

comonomer feed concentrations. Using this technology, in the 1990s BP

developed and patented a family of thermoplastic high polyacrylonitrile resins

called Amlon®. BP claims that for a given copolymer composition, the sequence

length of the AN sidegroups can be reduced sufficiently to enable melting.

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 61

When comparing the resultant copolymers from ‘scavenged’ and

traditional ‘flooded’ copolymerization reactions, the resulting structure of the

former would shift towards more regularly-alternating structures, because the

number of junctions between methyl acrylate and acrylonitrile junctions increases

and the distance between junctions is reduced. PAN crystallinity is generally

considered a result of the interactions and packing of acrylonitrile side groups, so

this less blocky structure would disrupt the pseudo-crystalline order more

frequently. Since the melting temperature of PAN copolymers has been shown

to be a function of the AN sequence length and if controlling the length is now

possible, a PAN copolymer could be made such that an optimum balance

between sequence length and melt temperature is achieved.

5.1 Experimental

5.1.1 Materials

The Institute of Textile Technology, at North Carolina State University,

Raleigh, NC now holds worldwide patent rights to Amlon® technologies. Amlon®

D resin is an 85% acrylonitrile : 15% methyl acrylate copolymer with a molecular

weight between 50k and 60k [78].

Dralon® type 251 bright 1.3 dtex of 93% acrylonitrile : 7% methyl acrylate

was supplied by Dralon GmbH, Dormagen, Germany. Exxon Mobil

polypropylene type 16050 was used to compare viscosities.

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 62

5.1.2 Solution 13C Nuclear Magnetic Resonance (NMR)

An Oxford Instruments 300 MHz for protons NMR spectrometer using 1H-
13
decoupling was used to collect C spectra. Type 507-HP round bottom NMR

sample tubes from Norell, Inc. were used. Approximately 50 mg of undried

Amlon® D and Dralon® fiber were dissolved in 0.6mL of deuterated d6-dimethyl-

sulfoxide. Spectra were recorded at 120º C and 20,000 scans were collected

[79].

The number average sequence length of acrylonitrile monomer units in

Amlon® D was determined from the experimentally observed comonomer triad

concentrations [80], using the following relation:

N101 + N001 + N000


n0 = [81],
1
N101 + N001
2

where n0 is the number average sequence length of acrylonitrile (A) units, and

N101 , N001, and N000 the concentrations of MAM (M = methacrylate unit), AAM

and MAA, and AAA comonomer triads, respectively.

Based on the actual mole fraction of acrylonitrile and methyl acrylate

comonomers in the feed f1 and f2, respectively, the mole fraction of comonomers

expected in the copolymer F1 and F2, respectively, were calculated using zero-

(Bernoullian) and first-order Markov or terminal model probabilities. First order

probabilities were calculated using the following equations:

r1f12 + f1f2
F1 = and F1 = 1− F2 [81]
r1f12 + 2f1f2 + r2 f22

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 63

where r1 and r2 are the monomer reactivity ratios of the AN and MA comonomers,

whose instantaneous values are 1.5 and 0.84 respectively. For AN and vinyl

acetate (VA), the ratios are 5.5 and 0.06, respectively [81].

5.1.3 Differential Scanning Calorimetry (DSC)

A Perkin-Elmer Diamond 7 DSC instrument calibrated with indium and

zinc was used to collect thermograms. Specimens were vacuum-dried for four

days and prepared by cutting pellets into granules with a razor blade. They were

weighed and encapsulated in volatile aluminum pans. A cyclic procedure was

used with the multiple melting at successive heating rates of 5, 10, and 20º

C·min-1 from 25 to 250º C [82].

5.2 Results and Discussion

5.2.1 Solution 13C Nuclear Magnetic Resonance

Most high acrylonitrile content, flooded free radical PAN homo- and

copolymers are known not to exhibit a melting endotherm at low ramping heating

rates [55, 83, 84]. Eby [71] proposed a method of predicting melting

temperatures for copolymers that includes comonomers as defects in their

crystalline regions, which Frushour [74, 85-87] demonstrated is generally

applicable to free radical polyacrylonitrile copolymers. Inference of the melting

point was made from wet melting points obtained in high pressure DSC scans

when mixing two parts water with one part copolymer. For many co- and

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 64

terpolymers of acrylonitrile, Eby’s model in association with Frushour’s wet

melting point depression accurately describe and predict the dry melting behavior

for some common comonomers.

Frushour’s model describing melting behavior for copolymers is

dependent on the comonomer concentration and the molar volume of their

sidegroups. Methyl acrylate and vinyl acetate have similar molar volumes, 36

and 39 cm3·mol-1, respectively. No dry-polymer melting point depression

constant is reported for PAN copolymers including MA, as “many of the polymers

decomposed before the melting could be measured.” Yet for VA, Frushour

reports a different wet melting point depression constant than that observed by

Slade et al., in dry melting experiments, the former of which is similar to the wet-

polymer melting point depression constant for MA [88]. Since a dry melting

temperature depression constant was observed for VA and not MA another

aspect of copolymer structure must affect their melting behavior.

The number average acrylonitrile sequence lengths of AN/MA and AN/VA

copolymers with various compositions were predicted and are compared with

those observed in Table 3 and Figure 9. Flooded free radical PAN comonomer

sequence lengths are predicted using zero (Bernoullian), first (Terminal) and

second (Penultimate) order Markov models from comonomer composition,

comonomer reactivity ratios, and observations of free radical AN/MA copolymers

[1] and AN/VA copolymers [89] by Brar, et al. Actual sequence lengths in

commercial high PAN fibers were calculated using triad concentrations obtained

by integration of the Amlon® and Dralon® 13


C NMR spectra recorded in d6-

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 65

dimethylsulfoxide at 120º C, such as presented in Figure 8. The second order

model shows the most accurate prediction for the free radical AN/MA up to 88%

acrylonitrile, see Figure 9. Yet, the observed sequence length of AN units in

Dralon® shows a nearly discontinuous shift/increase. The sequence length more

accurately reflects that predicted by random Bernoullian probability.

Furthermore, for Amlon® D the average length of contiguous AN units is

approximately thirteen, which is over two times the flooded equivalent and three

times less than in Dralon®. The resulting Amlon® D structures have block lengths

of intermediate size, whose specific size may not be possible in a flooded

copolymerization that yields a 85% AN: 15% MA copolymer. Thus, BP’s

scavenged reaction may offer sequence control beyond that of the flooded

monomer reactivity.

While the molar volume of the monomers MA and VA are similar, their

reactivity ratios with AN are significantly different. When comparing the predicted

and observed AN sequence lengths with respective comonomers, the VA

copolymer shows a more alternating and less blocky structure, which is

emphasized when comparing their comonomer triad concentrations. The

Penultimate model also more accurately describes the sequence lengths for

AN/VA. However, observed molar compositions are only 61, 64 and 75% AN.

The Terminal model is shown to predict a somewhat longer sequence length in

both MA and VA copolymers. When comparing the respective predicted AN

sequence lengths, the AN sequence length for the MA copolymer is over two

times larger than for the VA copolymer at the same AN molar composition. This

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 66

longer theoretical sequence length would provide longer ‘crystallizable’ AN units

and could explain the difference in melting behavior that is observed in these

copolymers. This would further address why shrinkage is more prevalent in

some acrylonitrile copolymers at higher comonomer concentrations [4, 90, 91].

Insertion of MA and VA units into PAN likely increases the conformational

flexibility of the copolymer relative to the PAN homopolymer [92, 93]. This

increase in conformational flexibility would be greater in AN/VA copolymers,

because the VA units are more randomly inserted as compared to the more

blocky insertion of MA units in Amlon® D. As a consequence, for the same

comonomer content AN/VA copolymers would be expected to exhibit greater

shrinkage than the Amlon® D copolymers.

Thermoplastic Polyacrylonitrile
67
Table 3. Comonomer Distributions and Number Average Sequence Lengths for Methyl Acrylate and Vinyl Acetate Copolymers
Polymer Molar Acrylonitrile Triad Comonomer Triad Number Number
Markov Feed Ratio
Comonomer Composition Concentrations Concentration Average Average X
Model
FAN FX fAN fX AAA AAX XAX XXX XXA AXA AN Length Length
0.71 0.29 0.50 0.41 0.08 0.08 0.41 0.50 3.45 1.41
0.85 0.15 0.72 0.26 0.02 0.02 0.26 0.72 6.67 1.18
0 0.88 0.12 0.77 0.21 0.01 0.01 0.21 0.77 8.33 1.14
0.93 0.07 0.86 0.13 0.00 0.00 0.13 0.86 14.29 1.08
0.98 0.02 0.94 0.06 0.00 0.00 0.06 0.94 33.67 1.03
0.71 0.29 0.62 0.38 0.40 0.49 0.10 0.13 0.44 0.43 2.87 1.55
0.85 0.15 0.79 0.21 0.63 0.34 0.03 0.04 0.30 0.67 5.06 1.23
1 Methyl 0.88 0.12 0.83 0.17 0.70 0.28 0.02 0.03 0.25 0.73 6.18 1.18
Acrylate 0.93 0.07 0.90 0.10 0.81 0.18 0.01 0.01 0.16 0.83 10.17 1.10
0.98 0.02 0.97 0.03 0.94 0.06 0.00 0.00 0.05 0.95 33.99 1.03
0.71 [1] 0.29 0.46 0.43 0.11 0.01 0.19 0.79 3.33 1.14
2
0.88 [1] 0.12 0.70 0.28 0.03 0.00 0.10 0.90 5.94 1.05
0.71 [1] 0.29 0.50 0.40 0.10 0.02 0.19 0.79 3.33 1.13
0.85 0.15 Amlon® 0.86 0.12 0.02 0.08 0.15 0.77 13.10 1.19
Observed
0.88 [1] 0.12 0.67 0.28 0.05 0.00 0.14 0.86 5.26 0.67
0.93 0.07 Dralon® 0.95 0.04 0.00 0.00 0.00 1.00 39.04 1.00
0.40 0.60 0.16 0.48 0.36 0.36 0.48 0.16 2.38 1.73
0.50 0.50 0.25 0.50 0.25 0.25 0.50 0.25 4.33 1.30
0 0.70 0.30 0.49 0.42 0.09 0.09 0.42 0.49 6.83 1.17
0.85 0.15 0.72 0.26 0.02 0.02 0.26 0.72 13.14 1.08
0.98 0.02 0.90 0.10 0.00 0.00 0.10 0.90 33.67 1.03
0.79 0.21 0.41 0.59 0.23 0.68 0.09 0.01 0.02 0.97 2.31 1.02
0.85 0.15 0.51 0.49 0.32 0.62 0.06 0.01 0.01 0.98 2.73 1.01
1 Vinyl 0.93 0.07 0.71 0.29 0.54 0.45 0.02 0.00 0.00 0.99 4.18 1.00
Acetate 0.98 0.02 0.90 0.10 0.82 0.18 0.00 0.00 0.00 1.00 10.71 1.00
[89] 0.40 0.60 0.14 0.46 0.41 0.00 0.08 0.92 1.58 1.04
2 [89] 0.50 0.50 0.22 0.49 0.29 0.00 0.05 0.95 1.87 1.03
[89] 0.70 0.30 0.45 0.43 0.11 0.00 0.98 0.95 3.05 1.34
0.61 [89] 0.39 0.40 0.60 0.15 0.47 0.38 0.00 0.08 0.92 1.63 1.04
Observed 0.64 [89] 0.36 0.50 0.50 0.20 0.50 0.30 0.00 0.05 0.95 1.82 1.03
0.75 [89] 0.25 0.70 0.30 0.45 0.43 0.12 0.00 0.98 0.95 2.97 1.34
Polymer Microstructural Analyses 68

Acrylonitrile
Triads
MAA AAA

MAM
DMSO-d6
Intensity

120.5 120 119.5 119 118.5 118

AMA MMA Methyl acrylate


Triads

174 173 172 171

MA AN
0
200 175 150 125 100 75 50 25 0
ppm

Figure 8. Solution 13C NMR Spectra of Amlon® D in d6-DMSO

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 69

Length
45

Length
40
Sequence
Sequence 35
30 Dralon®

Amlon® D
25
ANAN

Zero-order
20
Average

First-order
Average

15
Second-order
10
Number

5
Number

Observed [1]
0
0 .3 0 .4 0 .5 0 .6 0 .7 0 .8 0 .9 1
Mole Fraction of AN

Figure 9. Number Average Acrylonitrile Sequence Length for Methyl


Acrylate Copolymers by Mole Fraction of Acrylonitrile: Observed Lengths
and Zero-Order, First-Order, and Second-Order Markov Models Lengths

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 70

5.2.2 Differential Scanning Calorimetry

For Amlon® D at a ramp rate of 20º C·min-1 (Figure 11), the melting

thermograms show a small endothermic peak at approximately 218º C, which is

higher than observed for a free radical copolymer with the same molar

composition and occurs below the onset of thermal degradation. The

intermediate blocky structure in Amlon® D produces less stable paracrystals than

Dralon®, but more stable than those in equivalent AN/VA copolymers, as

evidenced by the DSC scans.

When considering Amlon® D and its equivalent free radical copolymer, the

larger Amlon® D crystals would normally be expected to be more stable and

require a higher melting temperature compared to free radical PAN copolymers.

However, the lack of a discrete melting endotherm likely means that the PAN

crystals in Amlon® D are ‘nonequilibrium’, where segmental dipole interactions

dominate the combination of related torsional, steric and/or angular constraints

on the copolymer chains. In other words, since the acrylonitrile sidegroups have

such strong dipole moments, the strength of their interchain interactions

supersedes the forces that usually govern the equilibrium conformations of a

regular vinyl chain. As such, the disruption of these Amlon® D ‘crystal’ units may

be a transient process and not a first-order thermodynamic transition, whereby

increased thermal energy gradually disrupts individual dipole attractions sufficient

for melt-processing below the onset of AN degradation. Since the melting

thermogram shows no discrete endotherm, the as-crystallized polymer likely

transitions from a more nonequilibrium crystal into a less nonequilibrium structure

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 71

with increased heating. Furthermore, at a heating ramp rate of 5º C·min-1

Amlon® D shows no glass transition or melting endotherm, which indicates that

the ‘218º C transition’ observed at higher heating rates likely occurs over a

discrete and measurable amount of time, ie., is not a first-order phase transition

typical of the melting of semi-crystalline polymers.

The DSC scans show a polymer with some order, although not

traditionally crystalline. Thermograms are compared for polymers dried for zero

and four days in Table 4 and Figure 11. The presence of water generally lowers

the glass transition and ‘melting’ temperatures. Drying the polymer widens the

range of the peaks and enthalpies of crystallization. At a rate of 5º C·min.-1, a

fusion peak is not discernable. Only at higher rates do heating peaks emerge,

however, their source is not defined and they are difficult to measure [94].

None of the thermograms show a typical equilibrium melting endotherm,

see Figure 11 a) and c). At 20º C·min-1 a slight fusion peak emerges. This

nondescript profile is similar to that recorded by Illers’ on a 85:15 AN:MA free

radical copolymer [95]. However, at 5º C·min-1 Amlon® D shows essentially no

change in the baseline behavior. Clear ‘recrystallization’ exotherms suggest,

however, that reordering is occurring, see Figure 11 b) and d). Similar changes

in the area of ‘recrystallization’ exotherm by ramp rate are observed for

poly(ethylene terephthalate) fiber, Figure 10. The change in exotherm peak for

PET becomes broader with increased ramp rate. For Amlon® D, the peaks

become narrower and larger magnitude. This further suggests that a different

type of crystal structure is formed than observed in semi-crystalline polymers.

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 72

5.2.3 Conclusions

The present investigation provides a microstructural basis for the

difference in bulk behavior between Amlon® D and flooded PAN copolymers,

because the acrylonitrile run length in Amlon® D is significantly different. This

intermediate intrachain microstructure supports crystals that are able to be

disrupted and “melt” at a lower temperature, rendering it melt processable. Most

importantly, this length is sufficient to impart mechanical stability by preventing

significant shrinkage upon heating. The Amlon® D paracrystalline domains “melt”

in a nonequillibrium fashion, and indicate that PAN copolymer paracrystals

generally conform in a mesomorphic manner that exhibit no equilibrium first-order

phase transition, unlike most semi-crystalline vinyl homo- and copolymers.

5 ºC·min-1
Heat Flow (mW)

10 ºC·min-1

20 ºC·min-1

125 150 175 200 225 250


Temperature (C)

Figure 10. Differential Scanning Calorimetry Cooling Thermograms for


Poly(ethylene terephthalate) Tire Cord Fibers

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 73

Table 4. Thermogram Inflection Points and Enthalpies


for Dried and Undried Amlon® D Pellets
Scan Temperature Enthalpy
State Peak
(ºC·m-1) (ºC) (J/g)
Glass 83.01
5 Melting none
Crystallization 189.58 -6.911
Glass 82.29
undried 10 Melting 213.65
Crystallization 174.63 -7.342
Glass 84.80
20 Melting 211.17
Crystallization 169.51 -7.914

Glass 85.43 -
5 Melting 215.47 -
Crystallization 187.63 -6.075
Glass 86.52 -
dried 10 Melting 218.22 -
Crystallization 179.50 -7.339
Glass 87.58 -
20 Melting 218.44 -
Crystallization 174.24 -8.195

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 74

.
.
20ºC·min-1
9

Heat Flow (mW)


7
10ºC·min-1
5
3
5ºC·min-1
1
.
.
.
50 75 100 125 150 175 200 225 250
Temperature (C)

. 5ºC·min-1
.
10ºC·min-1
Heat Flow (mW)

.
20ºC·min-1
.

.
.

150 160 170 180 190 200 210


Temperature (C)

Figure 11. Differential Scanning Calorimetry Thermograms for


Dried and Undried Amlon® D Pellet Thermograms

a) Dried melting.
b) Dried Crystallization.

Thermoplastic Polyacrylonitrile
Polymer Microstructural Analyses 75

.
.
20ºC·min-1
.

Heat Flow (mW)


.
. 10ºC·min-1
.
5ºC·min-1
.
.
.
.
50 75 100 125 150 175 200 225 250
Temperature (C)

.
5ºC·min-1
.

. 10ºC·min-1
Heat Flow (mW)

.
. 20ºC·min-1
.
.

.
150 160 170 180 190 200 210
Temperature (C)

Figure 11. Differential Scanning Calorimetry. (continued)

c) Undried Melting.
d) Undried Crystallization.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 76

6 Polymer Thermomechanical Analyses

The ability of a polymer to be converted from chip to product requires that

it maintain stability over a range of conditions. The rheological and mechanical

response to temperature characterizes its behavior. In order to extrude a

thermoplastic material, the optimal processing conditions must be identified.

Hygroscopic behavior in polymers is a function of the availability of

electronegative sidegroups that provide a site to attract water molecules.

Hydrophilic commodity thermoplastic polymers are routinely dried before

processing to improve the mechanical integrity through order and orientation.

Polyamide, with a regain of 3.5 – 4.5%, and polyethylene terephthalate, with a

regain of 0.3 – 0.4%, are the typical examples [44].

Flooded high polyacrylonitrile is one of the standard precursors for carbon

fibers, as it readily degrades before any resemblance of melting behavior is

observed. During the manufacture of carbon fiber from PAN, various byproducts

degas and weight is lost as the chain structure and sidegroups crosslink, cyclize

and convert to carbon [47, 96].

Due to the complexity and toxicity of related processing technologies, the

number of studies published regarding polyacrylonitrile rheological data is limited.

Most of the publications concern solutions and gels [97-102]. One recent paper

[103] reports data on dynamic viscosity for a range of acrylonitrile to methyl

acrylate concentrations for copolymers it claims to have synthesized in a manner

similar to the Amlon® patents by British Petroleum [76, 104-108].

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 77

6.1 Experimental

The overall objective of the thermomechanical studies was to develop an

understanding of the basic behavior of the polymer in molten form. Vacuum

drying to evacuate moisture and ensure consistent extrudate conditioned Amlon®

D was used. Isorate thermogravimetric analyses reported the regimes of phase

transition as a function of pyrolysis and a general range of temperatures suitable

for melt processing. Isothermal thermogravimetric analyses provided a

quantitative assessment of the polymer stability at different temperatures. The

tandem mass spectrometry reported the byproducts of degradation as an

indication of its mechanism. As capillary rheometry is the original method of

assessing the flow behavior of thermoplastic polymers and since the shear rates

are similar to those found in industrial processes, it is the chosen method of

evaluating the effects of melt time and temperature on viscosity [109].

6.1.1 Drying

Since the regain of polyacrylonitrile is 1-2% [44], Amlon® D was dried at

120ºC from one to four days at 4 mm of Hg.

6.1.2 Fourier Transform Infrared Spectroscopy (FTIR)

A Thermo-Nicolet Nexus 470 FT-IR ESP with Avatar Omni-Sampler

attachment from Thermo Electron Corporation was used for the FTIR scan. The

scan was corrected with attenuated total reflectance. A sample was prepared by

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 78

slicing a pellet with a razor blade to produce a 1mg sample that was placed on

the crystal. Thirty-two scans were collected.

6.1.3 Thermal Gravimetric Analysis (TGA) & Mass Spectrometry (MS)

A TA Instruments TGA2950 of New Castle, Delaware was utilized for TGA

with a tandem MS from Pfeiffer Vacuum Instruments Model GSD 301 T. A

Perkin-Elmer Pyris 1 TGA calibrated with nickel and zinc was also utilized for

studies for high temperature degradation. For isothermal scans, the system was

purged with Argon 1 hour before and 30 minutes after loading a sample. The

temperature was ramped from ambient conditions to 200°C at 20°C·min-1. The

system was equilibrated at the melt temperatures listed in Table 6. It was held

for 2 hours while observing weight loss and degassed ions.

High temperature degradation studies were conducted in the presence of

air and nitrogen at ramp rate of 20°C·min-1 from 50ºC to 950ºC [110].

6.1.4 Capillary Rheometry

A Haake Rheoflixer HT type 556-0100 single bore capillary rheometer

from Thermo Electron Corporation was used. Round die dimensions and shear

rates are found in Table 5 and the entrance angle was 90º.

The range of recommended spinning temperatures was obtained from

internal British Petroleum laboratory data. Polymer melt stability was determined

through thermal gravimetric analysis with mass spectrometry. Final

temperatures and melt times were chosen based on a consistent extrudate.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 79

These are shown in Table 6. Consistent polymer stream is one without defects

with similar viscosity curves. Collection methods were in accordance with ASTM

Standard D3835-02. Correction treatment and calculations based on standard

rheological procedures were performed.

Table 5. Capillary Rheometry: Die Dimensions and Shear Rates


Capillary
Length (mm) Diameter (mm) Shear Rate (s-1)
Die
1 10.0 2.0 10 – 2564
2 20.0 2.0 10 – 2564
3 30.0 2.0 10 – 2564

The Rheoflixer with mounted die was allowed to equilibrate at the desired

melt temperature. The single barrel was filled with dried polymer pellets. The

piston was attached and the head descended to contact the sample. Once a 12

bar pressure was reached, movement stopped and the polymer melt time began.

Once the melt time was reached, the sample was compressed for 60

seconds at 20 bar. Before extrusion, the pressure was decompressed to 10 bar

for 30 seconds. If the polymer flowed freely, shear rates began at 10 s-1 and

ended at 2564 s-1. Otherwise, if the polymer failed to exit the die at low shear

rates, shear sampling began at 2564 s-1 and ended at 10 s-1. Five data points

per shear rate 100 milliseconds apart were obtained once the piston pressure

stabilized at less than 2 bar·s-1.

The Bagley correction was calculated using pressure and apparent shear

rate, and L/R, where L is the die length and R is the die radius. True shear

stress was calculated using, σ R =


(∆P)R , where ∆P is the pressure change
2(L + e oR )

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 80

and eo is the end or Bagley correction [109]. The value was obtained as the y-

intercept of a linear fit between pressure and L/R.

Throughput Q was calculated using the relationship with apparent shear

4Q
rate γ& app = . The true shear rates were determined with the Rabinowitsch
πR 3

⎛ 3 1 dlnQ ⎞
procedure using the equation: γ& W = γ& app ⎜⎜ + ⋅ ⎟⎟ . True viscosity ηW was
⎝ 4 4 dlnσ R ⎠

σR
determined using η W = [109].
γ& W

Table 6. Extrusion Conditions: Melt Temperature and Melt Times

Material Temperature (ºC) Melt Times (min)


Amlon® D 210 25, 12, 9
Amlon® D 220 11, 5, 3
Amlon® D 230 7, 5, 3

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 81

6.2 Results and Discussion

6.2.1 Vacuum Drying

Batches of approximately 650 grams of polymer were vacuum dried.

Nearly 0.5% weight is lost after one day of drying, see Figure 12. On average

0.25% additional weight is lost during drying between days 2 and 4.

Given that the hydrophilic nature of these polymers is driven by hydrogen

bonding, water is both freely and weakly held in the bulk. The polar acrylonitrile

sidegroups in polyacrylonitrile provide sites of stronger attraction. Thus, it

requires additional time and energy to release the water bound there.

6.2.2 Chemical Composition

FTIR scans verify the contents of the polyacrylonitrile copolymerized with

methyl acrylate, see Figure 13. The 3350 cm-1 band shows residual water.

Methyl groups from the copolymer are indicated at 2920, 2850 and 1450 cm-1.

The acrylonitrile sidegroup is shown by the band at 2240 cm-1. Carbonyl groups

in methyl acrylate are shown by the 1730, 1650 and 1580 cm-1 bands. Thiols in

the mercaptan as well as ester links in the copolymer and/or maleate ester

stabilizer are indicated in peaks at 1230, 1170 and 1120 cm-1. Possible sulfate

residues from the initiator and surfactant are indicated with the 1120 cm-1 peak.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 82

0.75

Weight Loss (%)


0.5

0.25

0 1 2 3 4

Day

Figure 12. Weight Loss during Vacuum Drying of Amlon® D Pellets

0.0125
2920

0.01
1650
1580
1730

1450
3350
Absorbance

0.0075
2850

1170
1120
1230

0.005
2240

0.0025

0
4000 3000 2000 1000
Wavelength (1/cm)

Figure 13. Fourier Transform Infrared Spectra of Amlon® D Pellets

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 83

6.2.3 Thermal Stability

The ability of the polymer to maintain stability for prolonged periods of time

under isothermal conditions is a requisite for commercial fiber spinning. Figure

14 shows the isothermal weight loss for Amlon® D vacuum dried 1 day at 120 ºC.

Higher temperatures show larger amounts of weight loss. The amount of loss is

approximately 0.50, 1.00, and 1.50 % for 210, 220 and 230 ºC, respectively. The

weight loss for the 220 ºC scan shows an inconsistency with the others. This

could be caused by material variability in the batch polymerization process.

Amlon® D pellets are clear and yellow, which is due to the typical yellow

color of acrylonitrile monomer. The color gradually changes to shades of orange

during heating as it crosslinks and blackens as it pyrolizes. After isothermal

treatment, the specimens exhibited different shapes and colors. The 210 ºC

sample remained clear and in pellet shape, but turned light orange. The 220 ºC

specimen formed an opaque amber colored amoeba shape with large bubbles.

The 230 ºC scan resulted in a blackened solid also having large bubbles. The

shape and color changes indicate varying degrees of melting and degradation.

The degradation stages for Amlon® D under nitrogen and air purge is

shown in Figure 15. For both scans, the onset inflection occurs at approximately

400 ºC. A second inflection under both purges occurs at between 500 and 525

ºC. The final weight is approximately 42% of the original. Under nitrogen

Amlon® D stabilizes at a lower weight, while in air no stabilization is reached.

Final weight is approximately 30% of the original. Oxidation reactions drive the

degradation. Thus, degradation of Amlon® D is sensitive to oxygen [111].

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 84

100

99.8

Weight fraction (%)


99.6
99.4

99.2
99
98.8
98.6

98.4
40 60 80 100 120 140 160 180 200 220 240
Temperature (C)

Figure 14. Quasi-Isothermal Stability Thermogravimetric Analyses


of Undried Amlon® Pellets.

100
Weight Fraction (%)

75

air

50
nitrogen

25
0 250 500 750 1000
Temperature (C)

Figure 15. Isorate Degradation Thermogravimetric Analyses


of Undried Amlon® Pellets.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 85

6.2.4 Degradation Byproducts

The mass spectrometry data shows the products of degradation over the

duration of the scans, see Figure 16. Data is presented with mass fraction on the

x-axis as the scans were taken at isothermal temperatures for 2 hours

concurrently with TGA. Ions with maximum relative intensity measured in mass

over charge (m/e) above 1% include 17, 18, 20, 28, and 40. Tests for unequal

variances showed only ion 28 with a statistically similar data spread.

The composition of the polyacrylonitrile copolymerized with methyl

acrylate support ion composition including carbon, hydrogen, oxygen and

nitrogen. These are likely to be present in the form of water as well as monomer

and oligomer from the polymer and copolymer. Ion 17 corresponds to OH and

NH3. Ion 18 indicates H2O and NH4. CH2CN is shown by ion 20. Fragment 28

includes C2H4, CO, N2 and HCNH. Finally, ion 40 reflects C3H4 and CH2CN.

Changes in the trends of relative intensity are similar for ions 17 and 18.

Intensities for these curves are also comparable. As the material was dried only

for one day the loss of water is consistent with additional weight loss upon drying

for additional days.

Ions 20 and 40 also have similar trends in the change of relative intensity.

The scan at 230ºC shows consistently high emissions with no clear equilibrium

reached. Temperatures 210ºC and 220ºC show smaller intensities, then taper off

quicker. With an inflection point near the end of the scan, the lower two

temperatures reach a quasi-equilibrium state.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 86

Interestingly, the 230ºC scan shows lower amounts of ions 17 and 18 and

larger sustained amounts of ions 20 and 40. At temperatures 210 and 220ºC, ion

20 begins with the same intensity and tapers off gradually. A similar trend is

observed for ion 40. However, for the latter ion, all three scans originate with the

maximum intensity [112].

Given the chemical composition of the acrylonitrile co-methyl acrylate

copolymer, the ions detected during degradation at the three melt temperatures

are expected. The source of these ions as it relates to the stability of the

molecular weight cannot be determined. Size exclusion chromatography of

samples heated for long periods of time will indicate the mechanism of

degradation, whether detrimental chain scission or a negligible lowering through

a removal of end units or unreacted comonomers. In general, stability of Amlon®

D over prolonged exposures to elevated temperatures becomes a factor in a

processing operation only when the supply of molten polymer fails to be

delivered in a consistent manner. In other words, if not dealt with properly, zones

of low or no flow rate may lead to defective or low performance extrudate.

6.2.5 Melt Time

Limits to melt (dwell) time were determined empirically. Isothermal TGA

analyses provided insight into the relative weight loss by temperature. The lower

limit was determined once evidence of melt fracture disappeared [109].

Sharkskin was prevalent at low relative dwell times and high shear rates, see

Figure 17.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 87

Upper limits were determined from air bubbles. Different sizes and

placements of bubbles were observed. Large air bubbles were observed if

insufficient air evacuation or packing occurred. These are shown across the

entire melt diameter, wholly within the melt and on the surface of the melt. For

consistent, defect-free extrudate, care must be taken to properly pack pellets in

the capillary bore. Small bubbles are evident in insufficiently dried polymer for

polymers treated over exceedingly long dwell times. These bubbles are caused

by water vapor or gaseous products of degradation.

0.05

0.04
m/e 17; NH3,OH

0.03

0.02

0.01
98.5 98.75 99 99.25 99.5 99.75 100
Mass Fraction (%)

210ºC
220ºC
230ºC

Figure 16. Mass Spectrometry Fragments of Undried Amlon® D Pellets


Collected with Tandem Isothermal Thermogravimetric Analyses.

a) m/e 17.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 88

0.175

0.15

m/e 18; H2O


0.125

0.1

0.075

0.05
98.5 98.75 99 99.25 99.5 99.75 100
Mass Fraction (%)

0.23

0.22

0.21
CH2CN

0.2

0.19

0.18
98.5 98.75 99 99.25 99.5 99.75 100
Mass Fraction (%)

210ºC
220ºC
230ºC

Figure 16. Mass Spectrometry Fragments. (continued)

b) m/e 18.
b) m/e 20.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 89

0.025

0.02

N2,C2H4,CO,HCNH
0.015

0.01

0.005
98.5 98.75 99 99.25 99.5 99.75 100
Mass Fraction (%)

0.95
C3H4,CH2CN

0.9
210ºC
220ºC
0.85 230ºC

0.8
98.5 98.75 99 99.25 99.5 99.75 100
Mass Fraction (%)

210ºC
220ºC
230ºC

Figure 16. Mass Spectrometry Fragments. (continued)

c) m/e 28.
d) m/e 40.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 90

6.2.6 Capillary Rheometry

The polymer was first run undried at 220ºC, shown in Figure 18b. As

expected, water acts as a lubricant and reduces the viscosity. However, the

values of the apparent viscosity fluctuated. Small and large bubbles were readily

visible in the extrudate, Figure 17.

Consistent extrudate was obtained with polymer that was vacuum dried for

two days. The temperature effect on the viscosity of Amlon® D is clearly

illustrated in Figure 18a through 17c. True viscosity and true shear rate is

calculated by the ratio of die length to diameter using data from all dwell times.

The curves show nearly identical agreement between dwell times and L/D.

Viscosity decreases with increase in melt temperature and shear rate. The flow

supports the power law showing the relation between true shear stress and true

n
shear rate: σ R = m γ& R where m is consistency and n represents the power-law

index. Linear fit values are shown in Table 7.

Table 7. Power Law Index for Dried Amlon® D Capillary Rheometry

Temperature (ºC) Index (n)

210 0.3861

220 0.5122

230 0.6698

For comparison, polypropylene was run at 180ºC and its melt viscosity

behavior is shown in Figure 18d). At low shear rates, the viscosity is similar to

Amlon D at 230ºC. At high shear rates, the viscosity is lower.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 91

6.2.7 Conclusions

The basic thermal properties and rheological behavior of Amlon® D for

fiber spinning are reported. Prior to processing, drying was found essential. The

polymer is concluded to be pseudo-crystalline in bulk without defined glass or

melting peaks. However, crystallization, or reordering peaks are distinct. Quasi-

isothermal TGA analysis shows that the polymer is more stable at lower

temperatures. Isorate TGAs demonstrate that the polymer decomposes to a

stable carbonaceous material at high temperatures under inert conditions. Mass

spectrometry shows that the products of degradation are expected to be water,

monomers and oligomers. The rheological behavior of the molten polymer

follows the power law relation. A stable extrudate is observed under a range of

dwell times and temperatures. At high temperatures, the viscosity decreases

and the behavior is similar to polypropylene.

The results of the thermomechanical study provided the evidence to

support selection of a melt processing condition – 220ºC for a melt time of 5

minutes. This choice was made from the observation of stability as a function of

melt time. At 220ºC, the polymer could be melted relatively quickly with almost

negligible loss of weight.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 92

Figure 17. Melt fractures of Dried Amlon® D Extrudate


from Capillary Rheometer

a) 7.5X Sharkskin, 220ºC melt temperature and 3 minute dwell time.


b) 7.5X Bubbles, 230ºC melt temperature and 12 minute dwell time.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 93

1e5

1e4

True Viscosity (Pa s) 1e3

100

10
10 100 1e3 1e4 1e5
True Shear Rate (1/s)

1e5

1e4
True Viscosity (Pa s)

dried
1e3

undried
100

10
10 100 1e3 1e4 1e5
True Shear Rate (1/s)

Figure 18. True Viscosities by True Shear Rate for Dried and Undried
Amlon® D and Polypropylene by Temperature and Melt Time

a) at 210ºC.
b) at 220ºC.

Thermoplastic Polyacrylonitrile
Polymer Thermomechanical Analyses 94

1e5

1e4

True Viscosity (Pa s)


1e3

100

10
10 100 1e3 1e4 1e5
True Shear Rate (1/s)

1e5

1e4
True Viscosity (Pa s)

1e3

100

10
10 100 1e3 1e4 1e5
True Shear Rate (1/s)

Figure 18. True Viscosities by True Shear Rate. (continued)

c) at 230ºC.
d) Polypropylene at 180ºC.

Thermoplastic Polyacrylonitrile
Melt Extrusion 95

7 Melt Extrusion

The objective of the continuous fiber spinning study was to obtain an

indication of the achievable fiber properties and to identify potential optimal

parameters that should be investigated further at a later date. Fiber properties

produced by melt extrusion are generally dependent upon such physical

characteristics as crystallinity, molecular weight and chain orientation. The

following discussion regarding mechanical properties of fibers is confined to

semi-crystalline polymeric materials. In determining the fiber properties from a

given resin molecular weight can be considered as an independent variable, if

the polymer chains are stable during conversion to fibers. So chain orientation

becomes the most important extrusion factor in maximizing the fiber properties.

A number of continuous processing conditions allow orientation

enhancement. Some of the most prominent methods include a heated shroud,

step drawings, and steam drawing. A shroud heated to a temperature between

the glass and melting transition points extends the amount of time that the chains

have to disentangle and orient. A stepped drawing zone enables property

enhancements through stages by minimizing the tension placed on the fiber for

draw. Roll temperatures also play an important role in thermoplastic materials by

influencing the amount of plastic deformation, which is generally material

specific. Steam, especially for acrylic fibers, where paracrystallinity is driven by

the interactions between acrylonitrile sidegroups, imparts additional chain

Thermoplastic Polyacrylonitrile
Melt Extrusion 96

mobility through decoupling. A recent paper investigated the achievable Amlon®

fiber properties for stepped and steam drawing [113].

Draw ratio is often given as the ratio of draw roller and take-up speeds.

However, drawing also occurs between the spinneret and take-up roller.

Mechanical properties are thus analyzed at different stages of drawing – drawing

between the extruder and the winding roller (spin-draw ratio) and drawing

between the two heated draw rollers (draw ratio). Spin draw is affected by the

length of time the extrudate remains molten after exiting the die, which is in turn

dependent upon the rate of heat transfer. The point at which the molten polymer

stream solidifies is called the solidification point. During this phase, the change

in the radius can be significant, especially if the solidification point is delayed with

a heat shroud. The heat shroud slows the rate of heat loss from the fiber stream,

enables smaller diameter formation, and generally allows chains to orient more

fully along the axis of the fiber. A high amount of orientation is generally not

achieved through winding unless high speeds (> 2500m·min-1) are used.

Partially-oriented fibers are thus generally considered as the product of the

winding zone in melt extrusion [9].

Post spin-drawing imparts further chain orientation and mechanical fiber

enhancements through deformation due to an elongation that induces a stress.

True stress is defined as the ratio of force to final cross-sectional area.

Conventional stress, used to characterize fibers, is the ratio of force to original

cross-sectional area. The three main types of deformation are: 1) plastic fracture

where stress plateaus and breaks, 2) deformation strengthening and 3) neck-

Thermoplastic Polyacrylonitrile
Melt Extrusion 97

type deformation. Solution-spun acrylic fibers exhibit deformation strengthening

behavior before reaching a maximum draw ratio. With the aid of steam as a

lubricant, it can undergo a neck-type deformation [9]. Most melt-spun fibers

exhibit a neck-type deformation where the fiber yields at a specific stress and

maintains a constant value while it irreversibly elongates. At a point (sometimes

called the natural draw ratio), which is dependent upon the material, dimensions

and elongation rate, the fiber strengthens as it deforms. Fibers are generally

drawn through the natural draw ratio for end-use application.

Draw ratios are imparted with and without the assistance of an external

heat source. In ‘cold-drawing,’ co-existent drawn and undrawn sections that

correspond to variation in fiber diameter characterize the behavior. The extent of

draw ratio depends on the material, rate of elongation and molecular weight.

Major commodity fiber-forming thermoplastics, such as polypropylene, polyamide

and polyethylene terephthalate, behave in this way. The temperature at which

this process is conducted is usually below the glass transition. The elongation

exhibited in these types of materials can range between 100-1000%. In ‘heated-

drawing’, the cross-sectional morphology is uniform and homogeneous.

Temperatures well above glass transition are used to impart a plastic

deformation without a distinct yield. A different history of temperatures between

glass and melting transitions yield different results [114].

The type and magnitude of drawing have long been used as the

independent variables in characterizing the mechanical properties of fibers. The

effects of draw ratio in the two drawing zones – winding draw and draw ratio –

Thermoplastic Polyacrylonitrile
Melt Extrusion 98

will be analyzed. Draw ratio will be presented as a ratio of initial to final fiber

diameters and by a ratio of drawing roller speeds. The fiber diameter was

calculated using an estimation of density, based on the reported density of the

PAN homopolymer, 1.17-1.19 g·cm-3 [44]. As the density of the methyl acrylate

comonomer is 1.22 g·cm-3, an estimated density for Amlon® D of 1.20 g·cm-3 was

used. The total effective draw ratio will also be used to correlate properties and

is hence defined as,

extrudate radius
total effective draw ratio = .
final fiber radius

7.1 Experimental

7.1.1 Extrusion Device

A single bore melt extrusion device, shown in Figure 19, was used to

extrude a monofilament by varying throughput, windup speed and draw ratio. A

round die of 2.0mm diameter and length of 30.0 mm was used. A pressure

transducer was mounted on the piston base and a temperature transducer was

mounted in the top of the die.

The bore was loaded with pellets that had been dried for 4 days. The

pellets were melted for at least 5 minutes at 220ºC. After this time, the polymer

was forcefully compressed to evacuate air. If air bubbles were present in the

extrudate, a high occurrence of fiber breaks occurred which prevented collection

of a sample of sufficient size without defects. During the compression, the

Thermoplastic Polyacrylonitrile
Melt Extrusion 99

polymer exhibited stick-slip behavior resulting in release of audible noises. As

the piston descended, the pressure first increased to 1400 and 1500 kg·force and

then released or slipped. Afterwards, the pressure fell to between 1100 and

1200 kg·force. Then it built up to a threshold and released again and again. This

irregular cyclic buildup and release occurred with less severity under low piston

speeds. As the polymer began to melt, the piston lowered more regularly until a

constant pressure level was reached where stick-slip behavior was no longer

observed. After 5 minutes, the sample was decompressed to 1100 kg·force for

approximately 30 seconds. Then the piston speed was set to the desired

nominal throughput rates.

Often times, the polymer required an additional 20-30 minutes before the

pressure stabilized. The stick-slip behavior gradually diminished and finally

disappeared once the piston emptied at least half of the bore. This behavior is

attributed to the bore design, which melts the pellets in the final 2 inches where

the die is heated by the jacket. A frictionless coating inside the bore and/or pipes

may reduce this phenomenon. A stable pressure regime ranged between 1100

and 1250 kg·force depending on the piston speed.

The piston speeds, a direct measure of throughput, was varied at 1.25, 2.5

and 5 mm·min-1. Once the pressure stabilized, the fiber was collected at speeds

varying from 100 to 750 m·min-1 and drawn inline between two draw rollers

heated at 110 and 125ºC. In later trials, an external heat source set at 145ºC

installed just below the exit of the orifice was used to reduce the quenching

gradient and quench point. The fibers were collected in a bag using an aspirator.

Thermoplastic Polyacrylonitrile
Melt Extrusion 100

Fibers were successfully spun at all throughputs. However, larger flow

rates required higher-than-available winding and drawing speeds to achieve

deniers under 50. A piston speed of 1.25 mm·min-1 was chosen as the setting

because the maximum take-up speed was the limit of achievable spin draw

ratios. A set of winding and drawing speeds were chosen and samples were

collected for 5 minutes without a break. For a piston speed of 1.25 mm·min-1, the

settings are shown in Table 8.

Table 8. Melt Extrusion, Winding and Drawing Conditions


Shroud Winding Roll Drawing Roll Drawing Roll
Temperature Speed Speed Draw Ratio,
( ºC ) ( m·s-1 ) ( m·s-1 ) by Speed
0 100 100 0.00
0 100 200 1.00
0 100 300 2.00
0 100 400 3.00
145 100 100 0.00
145 100 200 1.00
145 100 300 2.00
145 100 400 3.00
145 200 200 0.00
145 200 300 0.50
145 200 400 1.00
145 200 500 1.50
145 200 600 2.00
145 300 300 0.00
145 300 400 0.33
145 300 500 0.66
145 400 400 0.00
145 400 500 0.25
145 400 600 0.50
145 500 500 0.00
145 500 600 0.20
145 600 600 0.00
145 600 700 0.16
145 700 700 0.00

Thermoplastic Polyacrylonitrile
Melt Extrusion 101

Crank

Speed
Piston
Pressure
Capillary
Temp.
Extruder

Aspirator
& bag
12’

Drawing roll
Drawing roll (heated)
(heated)
Winding roll
9”

Bore

Spinneret
2”

Heated jacket
4”

Shroud
Extrudate

Figure 19. Melt Extrusion Device with Detail of Bore and Shroud Setup

a) extruder and windup unit


b) capillary extruder detail

Thermoplastic Polyacrylonitrile
Melt Extrusion 102

7.1.2 Mechanical Properties

The fibers were allowed to condition overnight at 25ºC and 65% relative

humidity. The denier and mechanical properties were measured on fibers with

gauge length of 2.54 cm in accordance with ASTM Standard 3288 using a

Favimat, Textechno, Mödchengladbach, Germany, and a cross-head speed of

200 mm·min-1. Data for the force at break or tenacity and elongation at break for

ten fiber specimens were collected for each sample. Cross section shape and

diameter variation were recorded optically.

7.1.3 Size Exclusion Chromatography

A Waters 1525 binary high pressure liquid chromatography pump with a

Waters 2414 refractive index detector, Milford, Massachusetts, was used to

obtain chromatograms corresponding to the molecular weight of fiber samples.

The column was a single Polymer Labs PLgel, 5µm particle size, mixed-d pore

type, 300x7.5mm column with a flow rate of 1 ml·min-1. The column was heated

to 50ºC and the detector to 30ºC. A solution of 0.01M LiBr-N,N-dimethyl

formamide was mixed and used as the mobile phase [115].

Polytetrafluoroethylene syringe filters, solvent filters and coated vials were used

with all samples. Each sample was injected separately.

7.1.4 Wide Angle X-ray Diffraction

Wide angle x-ray diffraction scans were collected on a diffractometer

instrument using Cu Kα radiation from Omni Scientific Instruments, Inc. of Biloxi,

Thermoplastic Polyacrylonitrile
Melt Extrusion 103

MS. The voltage was set to 40 milli-volts and the current was set to 30 milli-

amperes. Reflection scans were collected from fibers approximately one-inch in

length between 2θ angles of 10 to 35º in steps of 0.1º with a count time of 5

seconds. Transmission patterns were collected from fibers of one-half-inch that

were mounted horizontally on Polaroid Type 57 film.

7.1.5 Differential Scanning Calorimetry (DSC)

A Perkin-Elmer Diamond 7 DSC instrument from EG&G, Wellesley, MA,

calibrated with indium and zinc, was used to collect thermograms. Fiber

specimens were conditioned at 25ºC and 65% relative humidity. Melt extrusion

conditions are given in Table 8. Approximately 2.5 – 3.5 mg of fiber was rolled

into a ball and sealed in volatile aluminum pans. All samples were successively

heated and cooled at 30ºC·min-1 in the scan range of 50 to 250ºC [82].

7.2 Results & Discussion

7.2.1 Melt Spinning

The throughput setting at 1.25 mm·min-1 provided a wide range of draw

ratios. These settings provided a basis for comparing the effect of an auxiliary

heat shroud and draw ratio on the tenacity and elongation at break. Using low

variation in the denier of the fibers as a criterion (Figure 20 and Table 9), the

extrusion was considered acceptable as the material behaved in a manner

similar to that of other thermoplastic fibers.

Thermoplastic Polyacrylonitrile
104

Table 9. Tenacity and Elongation for Meltspun Amlon® D Fibers by Total Effective Draw Ratio
Winding Drawing Mean Draw Draw
Shroud Spin
Roll Roll Total Ratio, Ratio, Linear Density Tenacity Elongation
Temperature Draw
Speed Speed Draw by by ( denier ) ( gf·den-1 ) (%)
( ºC ) Ratio
( m·s-1 ) ( m·s-1 ) Ratio Diameter Speed
0 100 100 5.35 5.35 0.00 0.00 26.18 ± 1.58 2.27 ± 0.22 31.96 ± 10.48
0 100 200 5.99 5.35 0.64 1.00 16.61 ± 0.43 3.83 ± 0.44 26.58 ± 2.37
0 100 300 6.38 5.35 1.03 2.00 12.93 ± 0.23 3.86 ± 0.47 22.03 ± 1.61
0 100 400 6.54 5.35 1.19 3.00 11.74 ± 0.75 4.07 ± 0.92 21.81 ± 2.22
145 100 100 5.31 5.31 0.00 0.00 26.98 ± 1.57 2.28 ± 0.20 58.56 ± 11.38
145 100 200 6.30 5.31 0.99 1.00 13.62 ± 1.59 4.26 ± 0.52 29.34 ± 3.42
145 100 300 6.75 5.31 1.44 2.00 10.29 ± 0.34 4.78 ± 0.62 23.96 ± 2.30
145 100 400 7.23 5.31 1.92 3.00 8.01 ± 1.80 5.80 ± 0.81 21.73 ± 2.74
145 200 200 5.94 5.94 0.00 0.00 17.21 ± 0.25 2.26 ± 0.21 31.71 ± 3.91
145 200 300 6.57 5.94 0.63 0.50 11.63 ± 1.74 4.02 ± 0.63 31.21 ± 4.88
145 200 400 6.78 5.94 0.84 1.00 9.66 ± 0.61 4.12 ± 0.75 24.49 ± 3.46
145 200 500 6.90 5.94 0.96 1.50 9.44 ± 0.22 4.36 ± 0.66 23.52 ± 1.89
145 200 600 7.64 5.94 1.70 2.00 6.28 ± 0.34 4.90 ± 0.92 20.67 ± 2.37
145 300 300 6.36 6.36 0.00 0.00 13.05 ± 0.24 2.24 ± 0.19 28.42 ± 3.88
145 300 400 7.16 6.36 0.80 0.33 8.16 ± 0.68 3.90 ± 0.33 22.91 ± 1.53
145 300 500 7.39 6.36 1.03 0.67 7.20 ± 0.81 4.38 ± 0.41 23.28 ± 1.91
145 400 400 6.95 6.95 0.00 0.00 9.20 ± 0.91 3.12 ± 0.24 21.09 ± 1.06
145 400 500 6.96 6.95 0.01 0.25 10.40 ± 0.22 3.05 ± 0.20 22.70 ± 1.91
145 400 600 7.24 6.95 0.29 0.50 7.83 ± 0.57 3.54 ± 0.44 21.33 ± 1.68
145 500 500 7.20 7.20 0.00 0.00 7.97 ± 0.26 3.34 ± 0.21 22.61 ± 1.66
145 500 600 7.32 7.20 0.12 0.20 7.48 ± 0.82 3.54 ± 0.42 20.53 ± 2.34
145 600 600 7.37 7.37 0.00 0.00 6.92 ± 1.65 3.57 ± 0.55 24.22 ± 7.11
145 600 700 7.43 7.37 0.06 0.17 7.05 ± 0.66 3.68 ± 0.41 22.49 ± 1.54
145 700 700 7.73 7.73 0.00 0.00 6.02 ± 0.61 3.64 ± 0.30 21.80 ± 1.20
Melt Extrusion 105

7.2.2 Mechanical Properties

Solution processed acrylic fiber is formed during a collapse of the spin

dope during the diffusion of the solvent [5, 7]. This creates structures having

scales and voids within the fiber itself as well as an irregular fiber surface. Weak

points through these defects are created that reduce the tenacity of the fiber.

Nearly all acrylic fiber commercially produced is therefore in a staple form [5,

116]. Melt processable fibers are widely know to have more regular surfaces and

diameters, which was verified for Amlon® [117]. With fewer defects, higher

tenacities are expected. Furthermore, while solution-spun fibers exhibit a

deformation strengthening response to elongation [9], fibers from Amlon® D with

a low spin draw ratio exhibited plastic deformation (Figure 21). Fibers with a

larger total draw ratio exhibited a deformation strengthening behavior, whose

load increases with elongation.

Fibers first wound without a heat shroud at 100 m·min-1 are considered the

control. Fibers extruded and wound at 100 m·min-1 with and without the shroud

showed similar linear densities and tenacities, Table 9. However, the elongation

of the fiber extruded with the heat shroud is almost twice as large. The shroud,

in this case, most likely acts to allow the molecular chains to reach higher

orientation as entanglements, generally considered to hinder the degree of

elongation and loosened up. Drawn to the same ratio, the fiber with the shroud

had equivalent elongation, lower linear density, and higher tenacity. While the

final total draw ratio of both fibers wound at 100 m·min-1 is similar, the tenacity of

Thermoplastic Polyacrylonitrile
Melt Extrusion 106

the fiber with the heat shroud was nearly twice as large as that of Dralon®, 5.80

gf·den-1 compared with 3.11 gf·den-1.

The tenacity of the fibers was investigated as a function the spin draw

ratio, draw ratio, and total effective draw ratio. Figure 22 shows the distribution

of tenacities for the total draw ratio. The scatter plot, in general, demonstrates an

increasing linear trend in tenacity with an increase in total draw ratio, with an

overall mean tenacity of 3.70 gf·den-1. Elongation in Figure 23 shows a

decreasing trend with increasing total draw ratio, with an overall average of 26%.

The two largest tenacities are shown by fibers which had the highest draw ratio,

or by fibers extruded with the heat shroud wound at 100 and 200 m·min-1 and

drawn to 400 and 600 m·min-1, respectively. Nearly all Amlon® D fibers showed

greater tenacities than Dralon®.

When examining fiber tenacities as a function of spin draw ratio or draw

ratio, it is clearly observed in Figure 24 and Table 9 that a larger draw ratio yields

higher tenacity fibers. The plot of tenacity against spin draw ratio shows a wide

range of tenacity values for fibers with the lowest winding speed, 100 m·min-1,

regardless of the draw ratio. As the winding speed increases, up to a speed of

700 m·min-1, the average tenacity decreases. Figure 24 illustrates that

regardless of whether a shroud is present, higher tenacities are achieved through

higher draw ratio. When examining tenacities as a function of draw ratios, the

fibers extruded with the aid of the shroud show a larger draw ratio, smaller denier

and higher tenacity than those without a shroud. Values of tenacity are spread

Thermoplastic Polyacrylonitrile
Melt Extrusion 107

more evenly when plotted against versus draw value calculated from the ratio of

diameters rather than roll speeds.

A similar trend in elongation behavior is observed when it is examined as

a function of spin draw ratio and draw ratio, see Figure 26 and Table 9.

Elongation and variation in it both decrease as the spin draw ratio increases. As

draw increases, regardless of whether a shroud is used, the elongation

decreases. Higher variation in elongation is seen in the fibers with the lowest

and the highest draw ratios. Low draw ratios would be expected to induce

variation in elongation if the polymer chains were disorderly arranged or had

paracrystallized in a locally-entangled manner. Highest draw ratios would most

likely induce variation in elongation from the amplified effects of diameter

inconsistencies in small denier fibers or high tensions that induce weak points.

7.2.3 Molecular Weight

The retention time of the samples is an indirect measure of molecular

weight and is useful in gauging relatively changes in molecular weights. In this

case, any changes in molecular weight during extrusion would be noted by large

changes in retention times (Table 10). Fiber samples were selected to

incorporate a range of processing conditions and draw ratios. No significant

change in retention time is observed for the samples. No substantial change in

molecular weight is observed, which indicates that the polymer is stable from

pellet to fiber in a simple melt extrusion device.

Thermoplastic Polyacrylonitrile
Melt Extrusion 108

Table 10. Size Exclusion Chromatography Retention Times for


Amlon® D Fibers
Shroud Winding Roll Drawing
Mean Total Retention
Temperature Speed Roll Speed
Draw Ratio Time (min)
( ºC ) ( m·s-1 ) ( m·s-1 )
0 100 100 5.35 5.689
0 100 400 6.54 5.736
145 100 100 5.31 5.737
145 100 400 7.23 5.714
145 200 500 6.90 5.737
145 300 300 6.36 5.720
145 300 400 7.16 5.728
145 400 400 6.95 5.725
145 400 500 6.96 5.731
145 500 500 7.20 5.747
145 600 600 7.37 5.743
145 700 700 7.73 5.750

Thermoplastic Polyacrylonitrile
Melt Extrusion 109

35

30

Linear Density (den)


25
Average Total
Effective Draw Ratio
20
5.31
15
5.35
10 5.94
5.99
5 6.3
5 5.5 6 6.5 7 7.5 8
6.36
Total Effective Draw Ratio
6.38
6.54
35 6.57
6.75
6.78
30 8
6.9
6.95
Linear Density (den)

6.96
25
7.16
7.2
20 7.23
7.24
7.32
15 7.37
7.39
7.43
10
7.64
7.73
5
5 5.5 6 6.5 7 7.5 8
Average Total Effective Draw Ratio

Figure 20. Linear Density (denier) of Melt-spun Amlon® D Fiber


by Average Total Effective Draw Ratio

a) by Total Effective Draw Ratio


b) by Average Total Effective Draw Ratio

Thermoplastic Polyacrylonitrile
Melt Extrusion 110

50

40

30

Load (gf)
20

10

0 10 20 30 40 50
Elongation (%)

50

40

30
Load (gf)

20

10

0 10 20 30 40 50
Elongation (%)

Figure 21. Load by Elongation Curves for Amlon® D Fiber, Plastic


Deformation and Deformation Strengthening

a) Shroud, Wind Speed = 100 m·min-1, Draw Speed = 100 m·min-1


b) Shroud, Wind Speed = 100 m·min-1, Draw Speed = 400 m·min-1

Thermoplastic Polyacrylonitrile
Melt Extrusion 111

6
Average Total

Tenacity (gf / den)


5 Effective Draw Ratio
5.31
4 5.35
5.94
3 5.99
6.3
2 6.36
6.38
5 5.5 6 6.5 7 7.5 8 6.54
Total Effective Draw Ratio 6.57
6.75
6.78
8
6.9
7
6.95
6.96
6
7.16
Tenacity (gf / den)

7.2
5 7.23
7.24
4 7.32
7.37
3 7.39
7.43
2 7.64
7.73
5 5.5 6 6.5 7 7.5 8
Average Total Effective Draw Ratio

Figure 22. Tenacity for Amlon® D Fiber by Total Effective Draw Ratio

a) by Total Effective Draw Ratio


b) Averaged by Extrusion Setting

Thermoplastic Polyacrylonitrile
Melt Extrusion 112

80

70

60 Average Total

Elongation (%)
Effective Draw Ratio
50
5.31
40 5.35
5.94
30 5.99
6.3
20 6.36
6.38
5 5.5 6 6.5 7 7.5 8 6.54
Total Effective Draw Ratio 6.57
6.75
6.78
8
7 6.9
6.95
6 6.96
7.16
Tenacity (gf / den)

5 7.2
7.23
4 7.24
7.32
7.37
3
7.39
7.43
2
7.64
7.73
5 5.5 6 6.5 7 7.5 8
Average Total Effective Draw Ratio

Figure 23. Elongations for Amlon® D Fiber


by Average Total Effective Draw Ratio.

a) by Total Effective Draw Ratio


b) Averaged by Extrusion Setting

Thermoplastic Polyacrylonitrile
Melt Extrusion 113

6
Average Total

Tenacity (gf / den)


5 Effective Draw Ratio
5.31
4
5.35
5.94
3
5.99
6.3
2 6.36
6.38
1 6.54
10 20 30 40 50 60 70 80
6.57
Elongation (%) 6.75
6.78
8
6.9
7
6.95
6.96
6
7.16
Tenacity (gf / den)

7.2
5
7.23
4 7.24
7.32
3 7.37
7.39
2 7.43
7.64
1 7.73
5 5.5 6 6.5 7 7.5 8
Spin Draw Ratio

Figure 24. Tenacity by Elongation and Spin Draw Ratio for Amlon® D Fiber
by Spin Draw Ratio.

a) by Elongation
b) by Spin Draw Ratio

Thermoplastic Polyacrylonitrile
Melt Extrusion 114

6
Average Total

Tenacity (gf / den)


5 Effective Draw Ratio
5.31
4
5.35
5.94
3
5.99
6.3
2
6.36
6.38
1
0 .25 .5 .75 1 1.25 1.5 1.75 2 6.54
6.57
Draw Ratio, by Diameter
6.75
6.78
8
6.9
7
6.95
6.96
6
7.16
Tenacity (gf / den)

7.2
5
7.23
7.24
4
7.32
7.37
3
7.39
7.43
2
7.64
1 7.73
0 .25 .5 .75 1 1.25 1.5 1.75 2
Draw Ratio, by Diameter

Figure 25. Tenacities by Draw Ratio for Amlon® D Fiber


by Diameter Draw and Speed Draw Ratios.

a) No shroud. Draw is a ratio of initial to final fiber diameter.


b) Shroud at 145ºC. Draw is a ratio of initial to final fiber diameter.

Thermoplastic Polyacrylonitrile
Melt Extrusion 115

6
Average Total

Tenacity (gf / den)


5 Effective Draw Ratio
5.31
4
5.35
5.94
3
5.99
6.3
2
6.36
6.38
1
0 .5 1 1.5 2 2.5 3 6.54
6.57
Draw Ratio, by Speed
6.75
6.78
8
6.9
7 6.95
6.96
6 7.16
7.2
Tenacity (gf / den)

5 7.23
7.24
4 7.32
7.37
3 7.39
7.43
2 7.64
7.73
1
0 .5 1 1.5 2 2.5 3
Draw Ratio, by Speed

Figure 25. Tenacities by Draw Ratio. (continued)

c) No shroud. Draw is a ratio of final to initial roller speeds.


d) Shroud at 145ºC. Draw is a ratio of final to initial roller speeds.

Thermoplastic Polyacrylonitrile
Melt Extrusion 116

80

70
Average Total
60 Effective Draw Ratio
Elongation (%) 50 5.31
5.35
40 5.94
5.99
30 6.3
6.36
20
6.38
10 6.54
5 5.5 6 6.5 7 7.5 8 6.57
Spin Draw Ratio 6.75
6.78
8
6.9
Figure 26. Elongations at Break for Amlon® D Fiber 6.95
6.96
7.16
by Spin Draw Ratio.
7.2
7.23
7.24
7.32
7.37
7.39
7.43
7.64
7.73

Thermoplastic Polyacrylonitrile
Melt Extrusion 117

80

70

60
Average Total

Elongation (%)
Effective Draw Ratio
50
5.31
40 5.35
5.94
30
5.99
20 6.3
6.36
10 6.38
0 .25 .5 .75 1 1.25 1.5 1.75 2 6.54
Draw Ratio, by Diameter 6.57
6.75
6.78
8
80 6.9
6.95
70 6.96
7.16
60
7.2
Elongation (%)

50 7.23
7.24
40 7.32
7.37
30 7.39
7.43
20
7.64
10 7.73
0 .25 .5 .75 1 1.25 1.5 1.75 2
Draw Ratio, by Diameter

Figure 27. Elongation at Break for Amlon® D Fiber


by Diameter and Speed Draw Ratio.

a) No shroud. Draw is a ratio of initial to final fiber diameter.


b) Shroud at 145ºC. Draw is a ratio of initial to final fiber diameter.

Thermoplastic Polyacrylonitrile
Melt Extrusion 118

80

70
Average Total
60
Effective Draw Ratio
Elongation (%) 50 5.31
5.35
40
5.94
30 5.99
6.3
20 6.36
6.38
10
0 .5 1 1.5 2 2.5 3 6.54
Draw Ratio, by Speed 6.57
6.75
6.78
8
6.9
6.95
80
6.96
70 7.16
7.2
60 7.23
Elongation (%)

7.24
50
7.32
40 7.37
7.39
30 7.43
20 7.64
7.73
10
0 .5 1 1.5 2 2.5 3
Draw Ratio, by Speed

Figure 27. Elongation at Break by Draw Ratio. (continued)

c) No shroud. Draw is a ratio of final to initial roller speeds.


d) Shroud at 145ºC. Draw is a ratio of final to initial roller speeds.

Thermoplastic Polyacrylonitrile
Melt Extrusion 119

7.2.4 Wide Angle X-ray Diffraction

7.2.4.1 Background

The anomalous nature of PAN homo- and co-polymers yields structures

that are dependent primarily upon preparation and processing and secondarily

upon the presence of any comonomer and its concentration. In nearly all wide

angle x-ray diffraction (WAXS) studies and even in cases where single crystals

were examined, the off-equatorial reflections or meridional scatterings are diffuse

at best. Correspondingly, Miller indices are often reported as hk0. As a result,

structural order is thought to be two dimensional. Based on WAXS, various

models of phase structure have been proposed [118], with the strongest

evidence suggesting a single phase [57] or combination of paracrystalline and

amorphous phases [119, 120].

Most analyses of WAXS diffraction patterns show a narrow peak at

2θ≈17º, a broad peak at 2θ≈27º and a small narrow peak at 2θ≈30º. The most

recent and refined [58] protocol for characterizing the degree of paracrystallinity

and amorphousness was developed by Gupta and Singhal [121]. The 2θ≈17º

peak was assigned paracrystalline, the 2θ≈27º peak amorphous and 2θ≈30º

paracrystalline. Based on the interplanar spacing and associated Miller indices,

In the Polymer Handbook [55], three crystalline unit cells are reported for PAN:

hexagonal, orthorhombic and tetragonal, corresponding to syndio- or iso-tactic

stereoregularity. Based on calculated and observed densities, the evidence that

Thermoplastic Polyacrylonitrile
Melt Extrusion 120

supports an orthorhombic unit cell, albeit a quasi-hexagonal form, depending on

the ratio of a and b axes, is more substantiated [65, 122-125].

Numerous attempts have been made to correlate stereoregularity with

structure but inconclusive evidence persists [65, 122-125]. The conformations

and structures appear to be dependent upon preparation and processing in

manners that are dissimilar to other semi-crystalline polymers. Most semi-

crystalline fiber forming polymers are understood in terms of discrete

stereoregularities such as atactic, isotactic or syndiotactic and their distinctive

packings. However, PAN has been shown to exhibit constant paracrystallinity

regardless of tacticity [62]. In 1961, Bohn, et al., suggested in a study of laterally

ordered polymers (polyacrylonitrile and poly(vinyl trifluoroacetate) that for PAN

the severity of the acrylonitrile dipole moments would render an irregular packing

even for a stereoregular PAN [57]. Yet, current studies continue to attempt to

relate properties with types of stereoregularity used to describe semi-crystalline

polymers [59-61]. However, efforts to classify structures in such a way may over-

simply and inaccurately describe the irregular conformations in solid state.

The chains for a PAN homo- or co-polymer are considered to conform in a

helical fashion [57, 126] that pack in rigid rods [127] of 6Å in diameter [57, 128]

having hexagonal or orthorhombic forms [125]. The pitch of the helix is

dependent upon the conformation of the acrylonitrile sidegroups [57]. From

powdered samples, Grobelny, et al., suggest that the degree of amorphousness

and paracrystallinity is not related to stereoregularity, but to a higher degree of

‘perfection.’ Perfection, in this case, is correlated to the smallest full-width at

Thermoplastic Polyacrylonitrile
Melt Extrusion 121

half-max (FWHM) measurements of the 2θ≈17º peak and larger paracrystallinity

coefficients as defined by Hinrichsen [129]. The more perfect structure is

suggested to prefer configurations [118]. In addition to tacticity, annealing was

shown to increase the paracrystalline perfection.

Inclusion of comonomers into the paracrystalline domain tends to reduce

the diameter of the PAN rods, but not necessarily their packing [130, 131]. Due

to an increase of chain flexibility from the absence of such strong dipole

moments in the comonomer, its presence depresses the glass transition

temperature, creates a lower helix pitch and prevents neighboring acrylonitrile

sidegroups from interacting. In so doing, the presence of comonomers lowers

the degree of perfection of the paracrystalline lattice. Yet depending on

temperature and time of annealing, perfection and narrowness of FWHM in

acrylonitrile copolymers can surpass that of the PAN homopolymer. The

rotational flexibility of the comonomer sidegroups further contributes to this

perfection as it facilitates more optimal pairing of acrylonitrile sidegroups [130].

7.2.4.2 Results and Discussion

Overall, the wide angle x-ray reflection spectra (Figure 28) and

transmission patterns (Figure 29) are relatively similar regardless of the

difference in total effective draw ratio. As the fibers with the heat shroud

imparted the highest tenacities, these were chosen for WAXS analysis. The

reflection spectra are consistent with those reported in the literature. A

Thermoplastic Polyacrylonitrile
Melt Extrusion 122

pronounced peak is observed at 2θ≈17º, a broad peak between 2θ angles of 25

and 29º and a third peak between 2θ angles of 29 to 30º. A new peak was

observed at 2θ≈32º for fibers with total effective draw ratios of 5.31 and 7.73,

which may show more distant reflections consistent with those found in highly

drawn fibers [126]. Interplanar d-spacing are also consistent with crystal studies

of homo- and co-polymer PANs. Additional scans will be needed to confirm this

peak and the paracrystal dimensions.

The degree of order or paracrystallinity was calculated based on the

technique described by Gupta and Singhal [121]. The degree of order is similar

to solution processed fiber heat treated at 110 and 150º in the same study. The

similarity of the degrees of paracrystallinity for the fibers prevents full conclusions

from being drawn. The mechanics of the helix deformation during drawing and

its effect on pairing of acrylonitrile sidegroups would likely affect these values. A

more detailed study of fiber x-ray diffraction and more precise control of peak

fitting should enable more precision in this calculation.

The FWHM of the 2θ≈17º peak suggest trends based on method of draw

and the total effective draw ratio. Grobelny found that the FWHM decreased with

paracrystals perfection. Comparing fibers with only spin draw ratio, the fiber with

total effective draw ratio of 7.73 shows a lower FWHM than the fiber with total

effective draw ratio of 5.31. Fibers drawn between the heated draw ratios

showed broader 2θ≈17º peaks. The values of which decreased with larger draw

ratio. More consistent values of FWHM for the 2θ≈27º and 2θ≈30º peaks will

Thermoplastic Polyacrylonitrile
Melt Extrusion 123

enable a discussion of the perfection of the paracrystals and degree of order

imparted through drawing.

The WAXS transmission patterns (Figure 29) show a more ordered

structure by virtue of the intensity of the equatorial reflections. Most patterns of

PAN fibers show more diffuse halos even for drawn fibers. In the case of the

Amlon® D fibers, some evidence of a halo is shown for the total draw ratio of

5.31. However, higher draw ratios show more defined equatorial reflections with

little diffuse meridional scattering. Similar to Colvin and Storr, with higher draw

Table 11. Wide Angle X-Ray Diffraction Reflection Spectrum Analysis of


Amlon® D Fibers
Para-
2 Theta d-spacing Area FWHM
Fiber crystallinity
(º) (Å) (%)
(%)
100 m·min-1 Wind 17.40 5.10 52.74 1.03
100 m·min-1 Draw 25.13 3.54 45.86 54.14 7.97
Total Draw Ratio = 5.31 29.85 2.99 1.40 0.92
400 m·min-1 Wind 14.43 5.08 59.55 1.49
400 m·min-1 Draw 25.66 3.47 37.19 62.81 7.87
Total Draw Ratio = 6.95 28.80 3.10 3.26 0.71
100 m·min-1 Wind 17.52 5.06 49.93 1.42
400 m·min-1 Draw 28.86 3.32 40.07 59.93 4.87
Total Draw Ratio = 7.23 30.37 2.94 10.00 0.88
700 m·min-1 Wind 17.32 5.12 52.75 0.92
700 m·min-1 Draw 27.42 3.25 42.32 57.68 8.29
Total Draw Ratio = 7.73 30.37 2.94 4.93 4.76

ratios through staged and steam drawing, a more 3-dimensional ordering may be

possible [126]. Since the historical method for calculation of paracrystallinity for

solution processed acrylic fibers is based on more diffuse meridional scattering, a

new method for melt processed acrylic fiber may be required.

In general, the plastic deformation made possible in this melt processable

high acrylonitrile copolymer may offer additional opportunities for optimization of

morphology and structure than observed in solution processed acrylic fibers.

Thermoplastic Polyacrylonitrile
Melt Extrusion 124

The relationship between the mechanics of chain deformation, dynamics of

intermolecular sidegroup interactions, perfection of the paracrystalline domains

and optimization of the mechanical properties can be studied both

computationally and empirically.

F5.31 Total Draw Ratio


F6.95 Total Draw Ratio
F7.23 Total Draw Ratio
F7.73 Total Draw Ratio
Count

10 15 20 25 30 35
2 Theta

Figure 28. Wide Angle X-Ray Diffraction Reflection Spectrum for Fibers
Extruded with a Heat Shroud.

Thermoplastic Polyacrylonitrile
Melt Extrusion 125

Figure 29. Wide Angle X-Ray Diffraction Transmission Patterns for Fibers
Extruded with a Heat Shroud

a) 100 m·min-1 Wind, 100 m·min-1 Draw


b) 400 m·min-1 Wind, 400 m·min-1 Draw

Thermoplastic Polyacrylonitrile
Melt Extrusion 126

Figure 29. Wide Angle X-Ray Diffraction Transmission Patterns.


(continued)

c) 100 m·min-1 Wind, 400 m·min-1 Draw


d) 700 m·min-1 Wind, 700 m·min-1 Draw

Thermoplastic Polyacrylonitrile
Melt Extrusion 127

7.2.5 Paracrystallinity

Regardless of whether a single paracrystalline domain pervades the

structure or a combination of amorphous and paracrystalline domain exist, the

enthalpy of ‘recrystallization’ or re-ordering in Amlon® D may be used to support

speculation about the effects of draw on sidegroup interactions. Observations by

Grobelny, et al. and others [118, 121, 130], suggested that the inclusion of

comonomers and processing affects the chain conformation and perfection of the

paracrystalline domain through its effects on optimal intrermolecular acrylonitrile

pairing. While the details of local segmental deformation and pairing are beyond

the scope of a macroscopic analytical technique like DSC, it is nevertheless a

useful tool to discuss how may be optimized through the chain conformation.

The degree of reordering, which hereon will be referred to as the degree

of paracrystallinity, is presented as a function of total effective draw ratio in Table

12 and Figure 30. More significant changes in enthalpy are shown for fibers

having a higher draw ratio and lower spin draw ratio. For the fibers extruded

without the shroud, the paracrystallinity reaches an optimum value, with the

highest total draw ratio imparted through draw between heated rollers, not

through spin draw, Figure 30b and 30c. The value of the optimum is the same as

that of the fiber extruded with the shroud at the same windup speed. Rather than

reaching an optimum, these fibers consistently exhibited a lower degree of

paracrystallinity with higher draw ratio. The fiber spun without the shroud at the

highest draw ratio showed an enthalpy similar to that given by the highest drawn

fiber extruded with the shroud. A similar trend is observed with fibers wound at

Thermoplastic Polyacrylonitrile
Melt Extrusion 128

200 m·min-1 with a shroud, Figure 30c. For fibers drawn with a shroud at 400

m·min-1 or more marginal changes in the enthalpy are observed. In general,

however, it is observed that enthalpy increases with draw, regardless of whether

it is spin draw or draw ratio. When comparing fibers with the same winding roll

speed, the highest draw ratios showed a decrease in the enthalpy.

Differences in enthalpy may be related to changes in intermolecular

bonding as affected by the amount of draw and rotational flexibility of

comonomer.

Table 12. Differential Scanning Calorimetry Data for Amlon® D Fiber


Winding Drawing Mean First First Second Second
Shroud
Roll Roll Total Reorder Reorder Reorder Reorder
Temperature
Speed Speed Draw Peak Enthalpy Peak Enthalpy
( ºC )
( m·s-1 ) ( m·s-1 ) Ratio ( ºC ) ( J/g ) ( ºC ) ( J/g )
0 100 100 5.35 162.25 8.356 162.05 8.331
0 100 200 5.99 161.89 8.589 161.41 8.649
0 100 300 6.38 163.83 9.082 163.54 8.996
0 100 400 6.54 161.93 7.466 161.85 7.742
145 100 100 5.31 179.57 9.235 179.67 9.093
145 100 200 6.30 177.19 8.427 177.23 8.356
145 100 300 6.75 178.65 7.867 179.75 7.755
145 100 400 7.23 161.44 7.451 161.05 7.424
145 200 200 5.94 171.39 7.604 178.06 7.816
145 200 300 6.57 164.26 8.747 164.20 8.594
145 200 400 6.78 163.36 8.506 163.01 8.561
145 200 500 6.90 179.68 9.689 180.21 9.254
145 200 600 7.64 162.80 8.143 162.94 8.150
145 300 300 6.36 167.37 8.796 167.74 8.826
145 300 400 7.16 163.90 8.887 163.84 8.816
145 300 500 7.39 169.92 9.433 172.01 9.515
145 400 400 6.95 160.40 8.219 160.64 8.294
145 400 500 6.96 162.67 8.564 162.44 8.564
145 400 600 7.24 163.71 8.928 163.52 9.016
145 500 500 7.20 163.83 8.824 163.80 8.841
145 500 600 7.32 169.99 9.094 170.45 9.108
145 600 600 7.37 177.33 8.075 179.38 8.258
145 600 700 7.43 176.78 8.498 176.78 8.723
145 700 700 7.73 165.73 9.471 165.78 9.379

Thermoplastic Polyacrylonitrile
Melt Extrusion 129

10

9.5
Total Effective
Draw Ratio
9

Enthalpy (J/g)
5.31
8.5 5.35
5.94
8 5.99
6.3
7.5 6.36
6.38
7 6.54
5 5.5 6 6.5 7 7.5 8 6.57
Total Effective Draw Ratio 6.75
6.78
6.9
10 6.95
6.96
9.5 7.16
7.2
9
Enthalpy (J/g)

7.23
7.24
8.5
7.32
7.37
8
7.39
7.43
7.5
7.64
7.73
7
5 5.5 6 6.5 7 7.5 8
Total Effective Draw Ratio

Figure 30. Second Cooling Enthalpies for Amlon® D Fiber


by Total Effective Draw Ratio.

a) For all fibers by average total effective draw ratio


b) No shroud. Windup Speed is 100 m·min-1.

Thermoplastic Polyacrylonitrile
Melt Extrusion 130

10

9.5
Total Effective
9 Draw Ratio

Enthalpy (J/g)
5.31
8.5 5.35
5.94
8 5.99
6.3
7.5 6.36
6.38
7 6.54
5 5.5 6 6.5 7 7.5 8
6.57
Total Effective Draw Ratio 6.75
6.78
6.9
10 6.95
6.96
9.5
7.16
7.2
9
Enthalpy (J/g)

7.23
7.24
8.5
7.32
7.37
8
7.39
7.5 7.43
7.64
7 7.73
5 5.5 6 6.5 7 7.5 8
Total Effective Draw Ratio

Figure 30. Second Cooling Enthalpies by Draw Ratio. (continued)

c) Shroud at 145ºC. Windup Speed is 100 m·min-1.


d) Shroud at 145ºC. Windup Speed is 200 m·min-1.

Thermoplastic Polyacrylonitrile
Melt Extrusion 131

7.2.6 Conclusions

Properties of Amlon® D fibers even though not yet optimized still show a

tenacity similar to that found in Dralon®. The tenacity of the Amlon® D fiber with

higher draw ratios should therefore be commercially acceptable. Smaller denier

fibers are currently being attempted through modification of the orientation zone

by incorporating a temperature device and drawing at different temperatures

[132].

Evidence of the possibility of additional property optimization beyond that

which is observed in solution processed acrylics is illustrated in wide angle x-ray

diffraction spectrum. Melt processability creates a structure that is able to

undergo chain deformation that affects intermolecular bonding, paracrystallinity

and the degree of perfection of the paracrystalline domain. Studies of fibers

drawn to higher ratios through staged and steam drawing as well as annealing

could further enhance the mechanical properties.

Thermoplastic Polyacrylonitrile
General Conclusions 132

8 General Conclusions

This study presented fundamental economical, technical and scientific

aspects of one version of the first innately thermoplastic and melt-processable

high polyacrylonitrile, Amlon® D. The following conclusions can be drawn from

those presented in corresponding sections of the thesis:

• The economic analysis presented shows that because of the low raw

material cost Amlon® will undoubtedly find increased consumption in the

world market. Mechanical property optimization due to melt processability

and intrinsic UV and chemical resistances will enable new applications.

• The intermediate length of acrylonitrile sidegroups is long enough for

mechanical integrity and dimensional stability while short enough to

become melt-processable. ‘Melting’ or de-coupling of sidegroups occurs

as a function of time and temperature. Cooling exhibits ‘recrystallization’

exotherms which can be considered sidegroup re-coupling.

• The rheological study shows that the behavior of molten polymer flow

varies little over a wide range of melt temperatures and times. Of the

processing parameters studied, the ideal temperature for melt processing

was found to be is 220ºC and melt time was found to be 5 minutes.

• Tenacity of Amlon® D fibers were on average higher than a commercial

solution spun fiber. Some extrusion conditions created fibers with

tenacities similar to most commercial poly(ethylene terephthalate) and

polyamides and one condition produced higher tenacities.

Thermoplastic Polyacrylonitrile
General Conclusions 133

8.1 Future Work

The science and engineering of melt extrusion of fibers is well developed.

As the basic thermodynamic and rheological properties of one type of Amlon®,

Amlon® D, are now established, optimization of fiber properties can now be

pursued. In particular, staged drawing, steam drawing and annealing

consistently improve fiber properties. High speed spinning may possibly further

enhance orientation and properties.

The Amlon® technology enables the synthesis of a range of resins

depending on comonomer, polymer structure and molecular weight. A

thermoplastic high polyacrylonitrile polymer has been an object of desire for

decades. Much of the longstanding mystery concerning the crystal morphology

of PAN is fortified by the lack of understanding their processability vis-à-vis

commodity thermoplastics. While this work has served to offer new knowledge

concerning the science of the polymer, much about this anomalous and

prodigious material remains to be understood. The ability to structure

acrylonitrile copolymers will create many opportunities to explore other

theoretical polymer science ramifications and to design polymers for specific

applications. Computational modeling will likely offer significant insight via the

speedy investigation of new polymers and the physical mechanisms that enable

enhanced properties. In general, as manufacturing technology progresses to the

point of precisely controlling the arrangement of matter and molecules, the

consequences of the structure-property relationship will continue to offer unique

insights into Nature, in all its breathtaking and inspiring forms.

Thermoplastic Polyacrylonitrile
Literature Cited 134

9 Literature Cited

1. Brar, A.S. and Sunita, Determination of Microstructure and Glass-

Transition Temperature of Acrylonitrile Methyl Acrylate Copolymers By


13
C-NMR Spectroscopy. Journal of Polymer Science, Part A, 1992. 30(12):

p. 2549-2557.

2. Horn, M.B., Acrylic Resins. 1960, New York: Reinhold Publishing

Company.

3. Sittig, M., Acrylic and Vinyl Fibers. 1972, Park Ridge, NJ: Noyes Data

Corporation.

4. Falkai, B.v., Dry-Spinning Technology, in Acrylic Fiber Technology and

Applications, J.C. Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

5. Frushour, B. and R. Knorr, Acrylic Fibers, in Handbook of Fiber Science,

M. Lewin and E.M. Pearce, Editors. 1985, Marcel Dekker: New York. p.

921.

6. Masson, J.C., Product Variants, in Acrylic Fiber Technology and

Applications, J.C. Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

7. Capone, G.J., Wet-Spinning Technology, in Acrylic Fiber Technology and

Applications, J.C. Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

8. Matzke, R.R., The Acrylic Fiber Industry Today, in Acrylic Fiber

Technology and Applications, J.C. Masson, Editor. 1995, Marcel Dekker,

Inc.: New York.

9. Ziabicki, A., Fundamentals of Fibre Formation. 1976, London: Wiley-

Interscience.

Thermoplastic Polyacrylonitrile
Literature Cited 135

10. Gadecki, F., History, in Acrylic Fiber Technology and Applications, J.C.

Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

11. Engelhardt, A., Worldwide and Regional Trends in Man-made Fiber

Production. Technical Textile Markets, 2002. 49: p. 58.

12. ---, Worldwide synthetic (noncellulosic) fiber production & producing

capacity by fiber except olefins: 1989 to 1995. Fiber Organon, 1994. 65(7):

p. 154-155.

13. ---, Worldwide olefin fiber production: 1989 to 1993. Fiber Organon, 1994.

65(7): p. 152-153.

14. ---, Worldwide synthetic (noncellulosic) fiber production & producing

capacity by fiber except olefins: 1990 to 1996. Fiber Organon, 1995. 66(7):

p. 134-135.

15. ---, Worldwide olefin fiber production: 1990 to 1994. Fiber Organon, 1995.

66(7): p. 136-137.

16. ---, World fiber production and producing capacity by fiber except olefin:

1995 to 2000. Fiber Organon, 1999. 70(7): p. 118-121.

17. ---, World olefin fiber production and producing capacity by product type:

1995 to 2000. Fiber Organon, 1999. 70(7): p. 126-127.

18. ---, World synthetic fiber production and producing capacity by fiber except

olefin: 1999 to 2004. Fiber Organon, 2003. 74(7): p. 130-133.

19. ---, World olefin fiber production and producing capacity by product type:

1999 to 2004. Fiber Organon, 2003. 74(7): p. 138-139.

20. ---, Sterling Chemicals, Inc. History. 2005, Hoover's Company Records.

Thermoplastic Polyacrylonitrile
Literature Cited 136

21. ---, Cytec Industries, Inc. History. 2005, Hoover's Company Records.

22. ---, Solutia, Inc. History. 2005, Hoover's Company Records.

23. Hajduk, F., Acrylic and Modacrylic Fibers, in Chemical Economics

Handbook Marketing Research Report. 2005, SRI Consulting: Menlo Park,

CA.

24. Sim, P.H., Solutia to Exit Acrylic Fiber Business, in Chemical Week. 2005.

p. 11.

25. Sharma, S., Role of acrylic fibre in the synthetic industry of India, in

Journal for Asia on Textile & Apparel. 2004. p. 38-39.

26. Lacson, J., Mono-, Di- and Triethylene Glycols, in Chemical Economics

Handbook Marketing Research Report. 2003, SRI Consulting: Menlo Park,

CA.

27. Lacson, J., Dimethyl Terephthalate (DMT) and Terephthalic Acid (TPA), in

Chemical Economics Handbook Marketing Research Report. 2004, SRI

Consulting: Menlo Park, CA.

28. Devenport, B., Y. Inoguchi, and S. Schlag, Polyolefin Fibers, in Chemical

Economics Handbook Marketing Research Report. 2004, SRI Consulting:

Menlo Park, CA.

29. Lacson, J., Propylene, in Chemical Economics Handbook Marketing

Research Report. 2004, SRI Consulting: Menlo Park, CA.

30. Hajduk, F., Nylon Fibers, in Chemical Economics Handbook Marketing

Research Report. 2002, SRI Consulting: Menlo Park, CA.

Thermoplastic Polyacrylonitrile
Literature Cited 137

31. Johnson, W.K. and G. Toki, Acrylonitrile, in Chemical Economics

Handbook Marketing Research Report. 2001, SRI Consulting: Menlo Park,

CA.

32. Chinn, H., W. Cox, and K. Yokose, Vinyl Acetate, in Chemical Economics

Handbook Marketing Research Report. 2004, SRI Consulting: Menlo Park,

CA.

33. Bizzari, S., Methyl Methacrylate, in Chemical Economics Handbook

Marketing Research Report. 2003, SRI Consulting: Menlo Park, CA.

34. Wade, B. and R. Knorr, Polymerization, in Acrylic Fiber Technology and

Applications, J.C. Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

35. Marash, S., et al., Polyester Fibers, in Chemical Economics Handbook

Marketing Research Report. 2004, SRI Consulting: Menlo Park, CA.

36. Ring, K.L., S. Schlag, and G. Toki, Unsaturated Polyester Resins, in

Chemical Economics Handbook Marketing Research Report. 2002, SRI

Consulting: Menlo Park, CA.

37. Borruso, A. and M.R. Devanney, Polypropylene Resins, in Chemical

Economics Handbook Marketing Research Report. 2004, SRI Consulting:

Menlo Park, CA.

38. Levy, J. and K. Yagi, Nylon Resins, in Chemical Economics Handbook

Marketing Research Report. 2004, SRI Consulting: Menlo Park, CA.

39. Linak, E., Adipic Acid, in Chemical Economics Handbook Marketing

Research Report. 2003, SRI Consulting: Menlo Park, CA.

Thermoplastic Polyacrylonitrile
Literature Cited 138

40. Linak, E., Caprolactam, in Chemical Economics Handbook Marketing

Research Report. 2003, SRI Consulting: Menlo Park, CA.

41. Malveda, M., Y. Inoguchi, and U. Loechner,

Hexamethylenediamine/Adiponitrile, in Chemical Economics Handbook

Marketing Research Report. 2004, SRI Consulting: Menlo Park, CA.

42. Lulay, A., Apparel End Uses, in Acrylic Fiber Technology and Applications,

J.C. Masson, Editor. 1995, Marcel Dekker, Inc.: New York.

43. Davenport, B., Y. Inoguchi, and S. Schlag, Polyolefin Fibers, in Chemical

Economics Handbook Marketing Research Report. 2004, SRI Consulting:

Menlo Park, CA.

44. Herlinger, K.H. and F. Schultze-Gebhardt, Ullmann's Encyclopedia of

Industrial Chemistry, in alphabetically arranged articles, W. Gerhartz, et

al., Editors. 1987, VCH Verlagsgesellschaft mbH: Weinheim.

45. Fourne, F., Synthetic Fibers. 1999, Munich: Hanser Publishers.

46. Butler, I., ed. The Filtration Technology Handbook. ed. L. Bergmann, E.

Homonoff, and G. Weismantel. 2000, INDA, Assocation of the Nonwoven

Fabrics Industry: Cary, NC.

47. Bahl, O.P., et al., Manufacture of Carbon Fibers, in Carbon Fibers, J.B.

Donnet, et al., Editors. 1998, Marcel Dekker, Inc.: New York.

48. Paiva, M.C., et al., UV stabilization route for melt-processible PAN-based

carbon fibers. Carbon, 2003. 41: p. 1399-1409.

Thermoplastic Polyacrylonitrile
Literature Cited 139

49. Hull, D. and T.W. Clyne, An Introduction to Composite Materials. 2 ed.

Cambridge Solid State Science, ed. D.R. Clarke, S. Suresh, and I.M.

Ward. 1996, Cambridge: Cambridge University Press.

50. Murdie, N., et al., Carbon-Carbon Matrix Materials, in Carbon-Carbon

Materials and Composites, J.D. Buckley and D.D. Edie, Editors. 2001,

National Aeronautics and Space Administration: Washington, DC.

51. Butler, I., Nonwovens Fabrics Handbook, ed. B. Subhash, et al. 1999,

Cary, NC: INDA, Association of the Nonwovens Fabrics Industry.

52. Butler, I., ed. The Spunbonded and Melt Blown Technology Handbook. ed.

E. Vaughn and L. Wadsworth. 1999, INDA, Assocation of the Nonwoven

Fabrics Industry: Cary, NC.

53. Baker, R.W., Membrane Technology and Applications. 2 ed. 2004, West

Sussex: John Wiley & Sons, Ltd.

54. Hoogers, G., Fuel Cell Components and Their Impact on Performance, in

Fuel cell technology handbook, G. Hoogers, Editor. 2003, CRC Press:

Boca Raton, FL.

55. Korte, S., Physical Constants of Poly(acrylonitrile), in Polymer Handbook,

J. Brandrup, Emmergut, E.H., Grulke, E.A., Editor. 1999, John Wiley: New

York. p. V-59.

56. Mandelkern, L., Crystallization of Polymers. 2 ed. Vol. 1. 2002,

Cambridge: Cambridge University Press.

Thermoplastic Polyacrylonitrile
Literature Cited 140

57. Bohn, C.R., J.R. Schaefgen, and W.O. Statton, Laterally Ordered

Polymers: Polyacrylonitrile and Poly(vinyl Trifluoroacetate). Journal of

Polymer Science, 1961. 55: p. 531-549.

58. Frushour, B., Arcylic Polymer Characterization in the Solid State and in

Solution, in Acrylic Fiber Technology and Applications, J. Masson, Editor.

1995, Marcel Dekker: New York.

59. Bajaj, P., M. Padmanaban, and R.P. Gandhi, Configurational Sequence

Lengths in Polyacrylonitrile and Poly(Acrylonitrile-Co-Haloalkyl Acrylate

Methacrylate)S Determined By C-13 Nmr. Polymer, 1985. 26(3): p. 391-

396.

60. Schaefer, J., High-Resolution Pulsed Carbon-13 Nuclear Magnetic

Resonance Analysis of Polyacrylonitrile. Macromolecules, 1971. 4(1): p.

105-107.

61. Yamadera, R. and M. Murano, Studies On Tacticity of Polyacrylonitrile.I.

High-Resolution Nuclear Magnetic Resonance Spectra of Polyacrylonitrile.

Journal of Polymer Science, Part A, 1967. 5(1): p. 1059-&.

62. Minagawa, M., et al., An Anomalous Tacticity-Crystallinity Relationship: A

WAXD study of Stereoregular Isotactic (83-25%) Poly(acrylonitrile)

Powder Prepared by Urea Clathrate Polymerization. Macromolecules,

2001. 34(11): p. 3679-3683.

63. Lee, L.W. and R.A. Register, Hydrogenated Ring-Opened

Polynorboronene: A Highly Crystalline Atactic Polymer. Macromolecules,

2005. 38: p. 1216-1222.

Thermoplastic Polyacrylonitrile
Literature Cited 141

64. Natta, G., P. Ganis, and P. Corradini, Prediction of Conformation of Chain

in Crystalline State of Tactic Polymers. Journal of Polymer Science, 1962.

58(166): p. 1191-&.

65. Holland, V.F., et al., Crystal Structure and Morphology of Polyacrylonitrile

in Dilute Solution. Journal of Polymer Science, 1962. 62(173): p. 145-&.

66. Tonelli, A.E. and M. Srinivasarao, Polymers from the Inside Out. 2001,

New York: Wiley-Interscience.

67. Krigbaum, W.R. and N. Tokita, Melting Point Depression Study of

Polyacrylonitrile. Journal of Polymer Science, 1960. 43(142): p. 467-488.

68. Krigbaum, W.R., Estimating the Unperterbed Dimensions of Polymer

Molecules. Journal of Polymer Science, 1958. 28: p. 213-221.

69. Flory, P.J., Theory of Crystallization in Copolymers. Transactions of the

Faraday Society, 1958. 51: p. 848-857.

70. Flory, P.J., Thermodynamics of Crystallization in High Polymers 4. A

Theory of Crystalline States and Fusion in Polymers, Copolymers, and

their Mixtures with Diluents. Journal of Chemical Physics, 1949. 17(3): p.

223-240.

71. Eby, R.K., First-Order Transition Temperatures in Crystalline Polymers.

Journal of Applied Physics, 1963. 34(8): p. 2442-&.

72. Slade, P.E., The Melting of Polyacrylonitrile. Thermochimica Acta, 1970.

1: p. 459-463.

Thermoplastic Polyacrylonitrile
Literature Cited 142

73. Kulshreshtha, A.K., et al., Effect of Comonomer Content and Annealing

On Morphological-Changes in Acrylic Copolymers and Fibers.76. Journal

of Applied Polymer Science, 1986. 31(5): p. 1413-1424.

74. Frushour, B.G., Melting Behavior of Polyacrylonitrile Copolymers. Polymer

Bulletin, 1984. 11(4): p. 375-382.

75. Lewis, F.M., F.R. Mayo, and W.F. Hulse, Copolymerization.2. the

Copolymerization of Acrylonitrile, Methyl Methacrylate, Styrene and

Vinylidene Chloride. Journal of the American Chemical Society, 1945.

67(10): p. 1701-1705.

76. Smierciak, R.C., E. Wardlow, and B. Lawrence, Process for Making an

Acrylonitrile, Methacrylonitrile and Olefinically Unsaturated Monomers, in

5,602,222. 1997, The Standard Oil Company: U.S.

77. Dimitratos, J., et al., Pseudosteady States in Semicontinuous Emulsion

Copolymerization. Journal of Applied Polymer Science, 1990. 40(5-6): p.

1005-1021.

78. British Petroleum Technical Data.

79. Tonelli, A.E., NMR Spectroscopy and Polymer Microstructure: the

Conformational Connection. Methods in Stereochemical Analysis. 1989,

New York: VCH.

80. Randall, J., Polymer Sequence Determination. 1977, New York: Academic

Press.

81. Odian, G., Principles of Polymerization. 4 ed. 2004, Hoboken, NJ: Wiley-

Interscience.

Thermoplastic Polyacrylonitrile
Literature Cited 143

82. Kim, Y.C., W. Ahn, and C.Y. Kim, A Study on Multiple Melting of Isotactic

Polypropylene. Polymer Engineering and Science, 1997. 37(6): p. 1003-

1011.

83. Dunn, P. and B.C. Ennis, Thermal Analysis of Polyacrylonitrile. Part 1.

The Melting of Polyacrylonitrile. Journal of Applied Polymer Science,

1970. 14: p. 1795-1798.

84. Hinrichsen, v.G., Untersuchungen zum Schmelzen von Polyacrylnitril. Die

Angewandte Makromolekular Chemie, 1971. 285: p. 121-127.

85. Frushour, B.G., Melting and Structure of Polyacrylonitrile. Bulletin of the

American Physical Society, 1980. 25(3): p. 352-353.

86. Frushour, B.G., A New Thermal Analytical Technique For Acrylic

Polymers. Polymer Bulletin, 1981. 4(5): p. 305-314.

87. Frushour, B.G., Water As a Melting-Point Depressant For Acrylic

Polymers. Polymer Bulletin, 1982. 7(1): p. 1-8.

88. Frushour, B.G., Melting Behavior of Polyacrylonitrile Copolymers. Polymer

Bulletin, 1984. 11: p. 375-382.

89. Brar, A.S. and Sunita, Sequence Determination of Acrylonitrile-Vinyl

Acetate Copolymers Prepared by Emulsion Polymerization using 13C-

NMR Spectroscopy, in Polymer Science: Contemporary Themes, S.

Sivaram, Editor. 1991, Tata McGraw-Hill: New Delhi. p. 582-587.

90. Henrici-Olive, G., S. Olive, and B.G. Frushour, Shrinkage Force and

Molecular Orientation in Polyacrylonitrile Fibers. Makromolekulare

Chemie, 1986. 187: p. 1801-1806.

Thermoplastic Polyacrylonitrile
Literature Cited 144

91. Lauterberg, W., W. Kimmer, and R. Schmolke, Das Schrumpfverhalten

von Fasern aus Acrylnitril-Vinylazetat-Copolymeren. Deutsche

Textiltechnik, 1968. 18: p. 657-658.

92. Tonelli, A.E., Glass Transition Temperatures of Regularly Alternative

Acrylonitrile-Vinyl Acetate Copolymers. Macromolecules, 1974. 10(3): p.

716-717.

93. Tonelli, A.E., Possible Molecular Origin of Sequence Distribution-Glass

Transition Effects in Copolymers. Macromolecules, 1977. 7(5): p. 632-634.

94. Huston, H.E. and A. Patil, Analytical Characterization of Weather-

Degraded Polyacrylonitrile Fiber. Journal of Applied Polymer Science,

1992. 44: p. 1523-1529.

95. Illers, v.K.H., Kalorimetrische Untersuchungen an Polyacrylnitril. Die

Makromolekulare Chemie, 1969. 3074: p. 278-281.

96. Burlant, W.J. and J.L. Parsons, Pyrolysis of Polyacrylonitrile. Journal of

Polymer Science, 1956. XXII: p. 249-256.

97. Bercea, M., S. Ioan, and S. Morariu, Oligo- and Polyacrylonitrile in Dilute

Solution. Excluded Volume Effect. Polymer-Plastics Technology and

Engineering, 2004. 43(2): p. 477-490.

98. Fritzsche, P., Relation Between Viscosity Number and Molecular Weight

of Polyacrylonitrile. Faserforschung und Textiltechnik, 1968. 19(12): p.

559-63.

Thermoplastic Polyacrylonitrile
Literature Cited 145

99. Ignatova, A.I., E.A. Pakshver, and S.A. Semenova, Viscosity of

concentrated polyacrylonitrile solutions. Karbotsepnye Volokna, 1966: p.

98-107.

100. Misra, G.S. and P.K. Mukherjee, The Relation Between the Molecular

Weight and Intrinsic Viscosity of Polyacrylonitrile. Colloid and Polymer

Science, 1978. 256(10): p. 1027-9.

101. Nicotera, I., et al., Mechanical properties of PAN-based get electrolytes:

small-amplitude oscillatory shear study. Plastics Rubber and Composites,

2004. 33(2-3): p. 125-129.

102. Uglanova, G.G., et al., Effect of solvent nature on the viscosity of

polyacrylonitrile solutions. Vysokomolekulyarnye Soedineniya, Seriya B:

Kratkie Soobshcheniya, 1974. 16(12): p. 902-905.

103. Rangarajan, P., et al., Effect of Comonomers on Melt Processability of

Polyacrylonitrile. Journal of Applied Polymer Science, 2002. 85: p. 69-83.

104. Ball, L.E., M. Wu, and E. Wardlow, Stabilizers for High Nitrile

Multipolymers, in 5,714,535. 1998, The Standard Oil Company: U.S.

105. Jorkasky, I., R.J., et al., Melt Spun Acrylonitrile Olefinically Unsaturated

Fibers and a Process to Make Fibers, in 6,114,034. 2000, The Standard

Oil Company: U.S.

106. Jorkasky, R.J., G.S. Li, and E.S. Percec, Process of Making High Nitrile

Composite Filaments, in 5,902,530. 1999, The Standard Oil Company:

U.S.

Thermoplastic Polyacrylonitrile
Literature Cited 146

107. Jorkasky, R.J., G.S. Li, and E.S. Percec, Process of Making High Nitrile

Composite Filaments, in 6,120,896. 2000, The Standard Oil Company:

U.S.

108. Smierciak, R.C., E. Wardlow, and B. Lawrence, Process for Making a High

Nitrile Multipolymer Prepared from Acrylonitrile and Olefinically

Unsaturated Monomers, in 5,618,901. 1997, The Standard Oil Company:

U.S.

109. Carreau, P.J., D.C.R. De Kee, and R.P. Chhabra, Rheology of Polymeric

Systems. 1997, Cincinatti: Hanser/Gardner Publications.

110. Conesa, J.A. and R. Font, Polytetrafluoroethylene decomposition in air

and nitrogen. Polymer Engineering and Science, 2001. 41(12): p. 2137-

2147.

111. Peebles, L.H., et al., On the Exotherm of Polyacrylonitrile - Pyrolysis of the

Homopolymer Under Inert Conditions. Carbon, 1990. 28(5): p. 707-715.

112. Xue, T.J., M.A. McKinney, and C.A. Wilkie, The Thermal Degradation of

Polyacrylonitrile. Polymer Degradation and Stability, 1997. 58(1-2): p. 193-

202.

113. Wu, Z., et al., Structure and properties of Melt-spun High Acrylonitrile

Copolymer Fibers via Continuous Zone-Drawing and Zone-Annealing

Processes. Thermochimica Acta, 2003. 396: p. 87-96.

114. Ward, I.M., Structure and Properties of Oriented Polymers, I.M. Ward,

Editor. 1997, Chapman & Hall: London.

Thermoplastic Polyacrylonitrile
Literature Cited 147

115. Azuma, C., M.L. Dias, and E.B. Mano, Size exclusion behavior of

polymers in amide solvents II. Molecular weight determination of

acrylonitrile polymers in N,N-dimethylformamide. Polymer Bulletin, 1995.

34: p. 593-598.

116. Eyicakar, O.B., Acrylic Fibre Properties and Their Influence on Rotor

Spinning, in Department of Textile Industries. 1981, University of Leeds:

Leeds. p. 239.

117. Davidson, J.A., et al., Investigation of Molecular Orientation in Melt-Spun

High Acrylonitrile Fibers. Polymer, 2000. 41: p. 3357-3364.

118. Grobelny, J., M. Sokol, and E. Turska, A Study of Conformation,

Configuration and Phase Structure of Polyacrylonitrile and Their Mutual

Dependence by Means of WAXS and 1H BL-NMR. Polymer, 1984. 25: p.

1415-1418.

119. Imai, Y., et al., Preparation and Characterization of Amorphous

Polyacrylonitrile. Journal of Polymer Science, Part B, 1970. 8: p. 281-288.

120. Joh, Y., Amorphous Polyacrylonitrile: Synthesis and Characterization.

Journal of Polymer Science, Part A1, 1979. 17: p. 4051-4067.

121. Gupta, A.K. and R.P. Singhal, Effect of Copolymerization and Heat

Treatment on the Structure and X-Ray Diffraction of Polyacrylonitrile.

Journal of Polymer Science, Polymer Physics Edition, 1983. 21: p. 2243-

2262.

122. Holland, V.F., Crystalline Morphology in Polyacrylonitrile. Journal of

Polymer Science, 1960. XLIII(142): p. 572-574.

Thermoplastic Polyacrylonitrile
Literature Cited 148

123. Klement, J.J. and P.H. Geil, Growth and Drawing of Polyacrylonitrile

Crystals Grown from Solution. Journal of Polymer Science, Part A-2,

1968. 6: p. 1381-1399.

124. Patel, G.N. and R.D. Patel, Single Crystals of High Polymers by Film

Formation. Journal of Polymer Science, Part A-2, 1970. 8: p. 47-59.

125. Lindenmeyer, P.H. and R. Hosemann, Application of the Theory of

Paracrystals to the Crystal Structure Analysis of Polyacrylonitrile. Journal

of Applied Physics, 1963. 34(1): p. 42-45.

126. Colvin, B.G. and P. Storr, The Crystal Structure of Polyacrylonitrile.

European Polymer Journal, 1974. 10: p. 337-340.

127. Grobelny, J., P. Tekely, and E. Turska, A Broad-Line Nuclear Magnetic

Resonance Investigation of Polyacrylonitrile Phase Structure and Chain

Conformation. Polymer, 1981. 22(12): p. 1649-1654.

128. Henrici-Olive, G. and S. Olive, Molecular Interactions and Macroscopic

Properties of Polyacrylonitrile and Model Substances. Advances in

Polymer Science, 1979. 32: p. 128.

129. Hinrichsen, v.G., Structural Changes of Drawn Polyacrylonitrile During

Annealing. Journal of Polymer Science, Part C, 1972. 38: p. 303.

130. Grobelny, J., M. Sokol, and E. Turska, Nuclear Magnetic Resonance and

Wide Angle X-ray Scattering in Poly(acrylonitrile-co-methyl acrylate): 1.

The Influence of Comonomer on Structural Characteristics and Annealing

Behavior. Polymer, 1989. 30: p. 1187-1196.

Thermoplastic Polyacrylonitrile
Literature Cited 149

131. Sokol, M., J. Grobelny, and E. Turska, Nuclear Magnetic Resonance and

Wide Angle X-ray Scattering in Poly(acrylonitrile-co-methyl acrylate): 1.

The Influence of Comonomer on Swelling Behavior. Polymer, 1991. 32: p.

2161-2166.

132. Kiang, C., The Influence of Polymer Characterization and Melt Spinning

Conditions on the Production of Fine Denier Poly(Ethylene Terephtalate)

Fibers, in College of Textiles, Department of Textile Engineering,

Chemistry and Science. 1990, North Carolina State University: Raleigh,

NC.

Thermoplastic Polyacrylonitrile

Das könnte Ihnen auch gefallen